Microbiology Monographs Volume 13
Series Editor: Alexander Steinbüchel Münster, Germany
Microbiology Monographs Volumes published in the series
Inclusions in Prokaryotes Volume Editor: Jessup M. Shively Vol. 1, 2006
Microbial Linear Plasmids Volume Editors: Friedhelm Meinhardt and Roland Klassen Vol. 7, 2007
Complex Intracellular Structures in Prokaryotes Volume Editor: Jessup M. Shively Vol. 2, 2006
Prokaryotic Symbionts in Plants Volume Editor: Katharina Pawlowski Vol. 8, 2009
Magnetoreception and Magnetosomes in Bacteria Volume Editor: Dirk Schüler Vol. 3, 2007
Hydrogenosomes and Mitosomes: Mitochondria of Anaerobic Eukaryotes Volume Editor: Jan Tachezy Vol. 9, 2008
Predatory Prokaryotes – Biology, Ecology and Evolution Volume Editor: Edouard Jurkevitch Vol. 4, 2007
Uncultivated Microorganisms Volume Editor: Slava S. Epstein Vol. 10, 2009
Amino Acid Biosynthesis – Pathways, Regulation and Metabolic Engineering Volume Editor: Volker F. Wendisch Vol. 5, 2007 Molecular Microbiology of Heavy Metals Volume Editors: Dietrich H. Nies and Simon Silver Vol. 6, 2007
Microbial Megaplasmids Volume Editor: Edward Schwartz Vol. 11, 2009 Endosymbionts in Paramecium Volume Editor: Masahiro Fujishima Vol. 12, 2009
Bernd H. A. Rehm Editor
Alginates: Biology and Applications
Editor Professor Dr. Bernd H. A. Rehm Institute of Molecular BioSciences Massey University, Private Bag 11222 Palmerston North 4442 New Zealand e-mail:
[email protected]
Series Editor Professor Dr. Alexander Steinbüchel Institut für Molekulare Mikrobiologie und Biotechnologie Westfälische Wilhelms-Universität Corrensstraße 3 48149 Münster Germany e-mail:
[email protected]
ISSN 1862-5576 e-ISSN 1862-5584 ISBN 978-3-540-92678-8 e-ISBN 978-3-540-92679-5 DOI: 10.1007/978-3-540-92679-5 Springer Dordrecht Heidelberg London New York Library of Congress Control Number: 2008942072 © Springer-Verlag Berlin Heidelberg 2009 This work is subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965, in its current version, and permission for use must always be obtained from Springer. Violations are liable to prosecution under the German Copyright Law. The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. Cover design: SPi Publisher Services Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com)
Preface
Alginates are well established as food additives and as encapsulation agents in biotechnology. Commercial production from harvested brown seaweeds commenced in the early twentieth century. Alginates belong to exopolysaccharides and are non-repeating copolymers of b-d-mannuronic acid (M) and a-l-guluronic acid (G) which are linked by 1–4 glycosidic bonds. These comonomers can be arranged in blocks of continuous M-residues (M-blocks), G-residues (G-blocks) or alternating residues (MG-blocks). The comonomer composition and arrangement strongly impact on the alginate material properties, which range in nature from slimy and viscous solutions to pseudoplastic materials. Brown seaweeds and only the two Gram-negative bacterial genera Azotobacter and Pseudomonas are capable of alginate production. The bacterial alginates are characterized by acetylation of M-residues to a variable extent at positions O-2 and/or O-3. The degree of acetylation was also found to affect the material properties of the alginate. In Pseudomonas aeruginosa, for example, alginate is mainly composed of M-residues mediating viscous solution properties required for its function as a biofilm matrix polymer. However, in Azotobacter vinelandii a pseudoplastic alginate with a high G-residue content is produced when this bacterium forms a desiccation-resistant cyst under adverse environmental conditions. The mature cysts are surrounded by two capsulelike layers containing a high proportion of alginate to maintain structural integrity and resistance to desiccation. In brown algae alginate is produced as an intercellular gel matrix contributing to the mechanical stability of the plants. Current knowledge of the genetics, gene regulation and biosynthesis of alginate is mainly based on extensive studies implementing the two bacterial species A. vinelandii and P. aeruginosa. Seminal work by Govan et al. and Chakrabarty et al. in the early 1980s kick-started research directed towards understanding the molecular mechanisms underlying alginate biosynthesis (Darzins and Chakrabarty 1984; Fyfe and Govan 1980; Ohman and Chakrabarty 1981). Interestingly, although the biosynthesis gene cluster was identified almost 30 years ago, the function of some genes essential for alginate biosynthesis has still not been assigned. The biosynthesis pathway leading to the formation of the activated alginate precursor GDPmannuronic acid is well understood; however, the polymerization of GDP-mannuronic acid and the export of the resulting alginate are poorly understood. The enzymes (epimerase, transacetylase, lyase) which catalyse modification of the nascent v
vi
Preface
polymannuronate and hence strongly impact on alginate material properties have been studied in more detail. In particular, the epimerization process, i.e. the introduction of G-residues at the polymer level, is of great interest from an applied point of view. Alginate is currently used for a variety of industrial purposes, and the production was hitherto exclusively based on brown seaweeds. However, the possibility to engineer bacterial production strains capable of producing tailor-made alginates for medical applications especially has become increasingly attractive. This monograph provides an overview of the state of the art of alginate material properties, genetics and biosynthesis as well as applications of alginates and tailormade alginates in medicine, food and biotechnology. Palmerston North and Münster November 2008
Bernd H.A. Rehm Alexander Steinbüchel
References Darzins A, Chakrabarty AM (1984) Cloning of genes controlling alginate biosynthesis from a mucoid cystic fibrosis isolate of Pseudomonas aeruginosa. J Bacteriol 159:9–18 Fyfe JA, Govan JR (1980) Alginate synthesis in mucoid Pseudomonas aeruginosa: a chromosomal locus involved in control. J Gen Microbiol 119:443–450 Ohman DE, Chakrabarty AM (1981) Genetic mapping of chromosomal determinants for the production of the exopolysaccharide alginate in a Pseudomonas aeruginosa cystic fibrosis isolate. Infect Immun 33:142–14
Contents
Material Properties of Alginates................................................................... Ivan Donati and Sergio Paoletti Alginate Production: Precursor Biosynthesis, Polymerization and Secretion ....................................................................... Bernd H.A. Rehm
1
55
Bacterial System for Alginate Uptake and Degradation ............................ Wataru Hashimoto, Yukie Maruyama, Takafumi Itoh, Bunzo Mikami, and Kousaku Murata
73
Enzymatic Alginate Modification ................................................................. Helga Ertesvåg, Svein Valla, and Gudmund Skjåk-Bræk
95
Alginate Gene Regulation.............................................................................. Dennis E. Ohman
117
Role of Alginate in Bacterial Biofilms .......................................................... Larry J. Halverson
135
Microbial Production of Alginates: Physiology and Process Aspects ....................................................................................... Wael Sabra and An Ping Zeng
153
Alginate-Based Blends and Nano/Microbeads ............................................ Kamalesh Prasad and Jun-ichi Kadokawa
175
Applications of Alginates in Food ................................................................. Iain A. Brownlee, Chris J. Seal, Matthew Wilcox, Peter W. Dettmar, and Jeff P. Pearson
211
vii
viii
Contents
Alginate and Its Comonomer Mannuronic Acid: Medical Relevance as Drugs.......................................................................... Abbas Mirshafiey and Bernd H.A. Rehm
229
Index ................................................................................................................
261
Material Properties of Alginates Ivan Donati and Sergio Paoletti
Contents 1 2
Sources and Applications..................................................................................................... Chemical Structure and Sequence ....................................................................................... 2.1 Chemical Composition ............................................................................................... 2.2 Comonomer Sequence ................................................................................................ 2.3 Natural Substituents .................................................................................................... 2.4 Alginate as a Polyelectrolyte ...................................................................................... 2.5 Conformational Properties of the Alginate Chain ...................................................... 3 Molecular Weight and Stability ........................................................................................... 3.1 Molecular Weight and Molecular Weight Distribution of Native Alginates .............. 3.2 Chain Degradation ...................................................................................................... 3.3 Chemical Modifications of Alginate ........................................................................... 4 Alginate Solutions................................................................................................................ 4.1 Thermodynamic Properties ......................................................................................... 4.2 Nonequilibrium Properties .......................................................................................... 5 Ion-Binding Properties......................................................................................................... 6 Alginate Gels ....................................................................................................................... 6.1 Gel Formation Properties ............................................................................................ 6.2 Hydrogel Stability ....................................................................................................... 6.3 Mechanical Properties................................................................................................. 6.4 Diffusion Properties .................................................................................................... References ..................................................................................................................................
2 3 4 5 10 12 15 17 17 18 20 21 21 23 26 30 30 35 37 45 46
Abstract The present chapter deals with the description of the main characteristics of alginate as a material. Sources and the chemical structure of the polysaccharide are discussed. Particular attention is devoted to the definition of the comonomer sequence and its determination together with chemical and biochemical modifications of alginate. The stability of alginate molecules is discussed from the point of view of the effect on the polymer molecular weight. The characteristics of the polysaccharide in solution are tackled looking both at the equilibrium properties (thermodynamic I. Donati (), S. Paoletti Department of Life Sciences, University of Trieste, Trieste, 34127, Italy e-mail:
[email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_1, © Springer-Verlag Berlin Heidelberg 2009
1
2
I. Donati and S. Paoletti
properties) and at the nonequilibrium ones (viscosity in dilute and semidilute solution). The second part of this contribution is focused on the analyses of the ion-binding properties of alginate towards divalent ions. A general description of the “egg-box” model is provided together with a brief overview of recent findings on this topic. The mechanism of hydrogel formation and the description of the mechanical properties of alginate gels in small and large deformation ranges are addressed.
1
Sources and Applications
“Alginate” is a collective term for a family of polysaccharides produced by brown algae (Painter 1983) and bacteria (Gorin and Spencer 1966; Govan et al. 1981). Alginic acid was first discovered, extracted, and patented by Stanford (1881, 1883a, b). This polysaccharide was recognized as a structural component of marine brown algae (Phaeophyceae), where it constitutes up to 40% of the dry matter and occurs mainly in the intercellular mucilage and algal cell wall as an insoluble mixture of calcium, magnesium, potassium, and sodium salts (Haug and Smidsrød 1967). The presence of alginate provides the mechanical strength and flexibility of the seaweed and, additionally, acts as water reservoir preventing dehydration once part of the seaweed has been exposed to air. Alginate can thus be regarded as having the same morphophysiological properties in brown algae as those of cellulose and pectins in terrestrial plants. Several bacteria, such as Azotobacter vinelandii and numerous species of Pseudomonas, produce an exocellular polymeric material which resembles alginate (Gorin and Spencer 1966; Linker and Jones 1966; Govan et al. 1981). In fact, alginate is an essential part of the protective cyst of dormant Azotobacter species and is the main constituent of biofilms of both Azotobacter and Pseudomonas (Campos et al. 1996), where it protects the bacterium from the host immune system. Pseudomonas aeruginosa, an opportunistic pathogen which colonizes the lungs of patients affected by cystic fibrosis, is able to secrete a viscous slime containing notable amounts of alginate (Pedersen et al. 1990). Secretion of alginate by Pseudomonas syringae has been reported to be associated with plant infection and colonization (Yu et al. 1999; Keith et al. 2003). To date, all commercially available alginates have been extracted from brown algae, mainly from Laminaria hyperborea, Macrocystis pyrifera, Laminaria digitata, Ascophyllum nodosum, Laminaria japonica, Ecklonia maxima, Lessonia nigrescens, and Durvillaea antarctica. The composition of alginate in different seaweeds varies according to seasonal and growth conditions, as well as within different parts of the plant (Andresen et al. 1977; Indergaard and Skjåk-Bræk 1987). This variability in composition stems from the different requirements, in terms of mechanical properties, of the plants. For example, in L. hyperborea, which grows in very exposed coastal areas, very high mechanical rigidity is required in the stipe and in the holdfast, whereas high flexibility is needed in the leaves that float on streaming water. The worldwide annual industrial production of alginate is estimated to be 30,000 metric tons, which is probably less than 10% of the biosynthesized material in crops of macroalgae. These figures allow us to consider such a polysaccharide as
Material Properties of Alginates
3
an unlimited and renewable resource even for a steadily growing industry (Draget et al. 2005). In addition, alginate production by fermentation is also technically possible, although it does not meet the requirement of economic feasibility. The main industrial applications of alginate as a natural polymeric material are linked to its stabilizing, viscosifying, and gelling properties and its ability to retain water. Alginate is largely used as a viscosifier in textile printing because of its shear-thinning characteristics. Its use has also been reported to increase color yield and brightness. In addition, various applications of alginate in the food industry are currently being exploited. The ability of this polysaccharide to stabilize aqueous mixtures, dispersions, and emulsions together with its gel-forming and viscosifying properties represent key features for the use of alginate in food applications. Several alginate-based restructured products (pet food, reformed meat, onion rings, crabsticks, to name a few) are available on the market for large-scale distribution (Cottrell and Kovacs 1980; Littlecott 1982; Sime 1990). In addition, alginate is widely used as an additive for the production of low-sugar jam, jellies, and fruit fillings (Toft et al. 1986). In these applications, synergistic interactions with proteins and/or other polysaccharides are exploited. Alginates well meet all the requirements for their use in pharmaceutical and medical applications. They have been largely used in wound dressings, dental impression, and formulations for preventing gastric reflux. However, the most advanced biotechnological and biomedical application of alginate resides in its use as a hydrogel for cell immobilization for applications ranging from production of ethanol from yeast cells and of antibiotics or steroids (Smidsrød and Skjåk-Bræk 1990) to transplantation and cell therapy (Lim and Sun 1980; Winn et al. 1991; Soon-Shiong et al. 1994; Hasse et al. 2000; Joki et al. 2001; Rokstad et al. 2002). In the latter case, alginate gel is used as a selective immune barrier to protect the transplanted cells from the host immune system. There has been a notable increase in the number of alginate applications in recent years and the possibility of using such a polysaccharide for advanced biomedical therapies requires very detailed knowledge of its molecular characteristics. In fact, a clear understanding of the structure–function relationships is crucial for successful preparation of refined alginates for those applications where specific biochemical and physicochemical features have to be met. In addition, engineering of alginate molecules, by tailor-making their composition and properties or by introducing cell-specific signals, represents an important step forward for future novel applications in the biotechnology field.
2
Chemical Structure and Sequence
Alginate can be regarded as a paradigm of how strongly the macroscopic properties of the polymer depend on the subtle details of its chemical composition and on the ensuing physicochemical properties in well-defined environments. For this reason we decided to provide a critical overview of such structural aspects as a foundation for the behavior of alginate as a material.
4
2.1
I. Donati and S. Paoletti
Chemical Composition
Although the discovery of alginate can be traced back to the end of the nineteenth century, the complete elucidation of its chemical structure required several years. The presence of nitrogen in the polymeric structure of the polysaccharide, proposed by Stanford himself, was definitely ruled out by additional investigations, and the presence of uronic acids, specifically mannuronic acid, as a fundamental structural feature of the alginate backbone was found in 1926 (Atsuki and Tomoda 1926; Schimdt and Vocke 1926). The glycosidic bonds connecting C-1 of one mannuronic unit to C-4 of the following one were found to display a b configuration; hence, alginate was believed to be the brown seaweed counterpart of cellulose. However, it was later elucidated (Fisher and Dörfel 1955) that another uronic saccharide, i.e., a-l-guluronic acid (G), was constitutively present in the polysaccharide. This finding revealed the intrinsic complexity of the polysaccharide and boosted intense research on analytical methods for the characterization of the alginate chain from the point of view of the chemical composition. In fact, it soon became very clear that alginate displayed a marked chemical inhomogeneity and that the content of the two different uronic groups varied over a wide range depending on the algal source and the part of the plant considered. Moreover, Haug and coworkers, by means of partial acid hydrolysis and fractionation experiments, succeeded in revealing the blockcopolymer nature of alginate. In particular, three fractions were isolated: two of these contained almost exclusively α-l-guluronic acid (G) and b-d-mannuronic acid (M) residues, respectively, while the third one was composed of both uronic acids in almost equal proportion (Haug and Smidsrød 1965; Haug et al. 1966, 1967c). As such, alginates do not possess a regular repeating unit. They can be described as linear binary copolymers of 1–4-linked M and G residues arranged in a blockwise pattern with homopolymeric regions of G residues (G-blocks) and a homopolymeric region of M sequences (M-blocks) interspersed by regions in which the two groups coexist in a strictly alternating sequence (MG-blocks) (Fig. 1) (Haug et al. 1966). The saccharide units composing the alginate chain adopt different conformations, i.e., 4C1 and 1C4 in the case of mannuronic acid and guluronic acid, respectively. It follows that diequatorial linkages connect mannuronic acid residues in M-blocks which assume a flat ribbonlike chain conformation. In contrast, diaxially linked guluronic acid residues lead to a buckled and more rigid structure for the G-blocks. MG-blocks are characterized by alternating axial–equatorial and equatorial–axial glycosidic bonds connecting the residues; such variation in the “tacticity” of the bonds accounts for the greater flexibility of the MG-block with respect to the other sequences. As a matter of fact, the rigidity of the chain blocks was reported to decrease along the series GG>MM>MG (Smidsrød and Whittington 1969; Stokke et al. 1993). These results have been revisited recently (Vold et al. 2006) and the block components of natural alginates, when analyzed separately, seem to show essentially the same chain stiffness, irrespective of their chemical composition.
Material Properties of Alginates
5
a HO O
NaOOC HO
OH
HO
OH
OH NaOOC
O OH
HO ManA (M)
GulA (G)
b NaOOC
OH OH O
O
OH
NaOOC
HO O
NaOOC HO O
O O HO OH
NaOOC
O O HO
NaOOC
G G - block
HO O
O
OH
G
M
M M - block
HO
NaOOC
O O HO
G
M GM - block
Fig. 1 Alginate chemical structure. a The 4C1 conformation of b-d-mannuronic acid (M) sodium salt and the 1C4 conformation of a-l-guluronic acid (G) sodium salt. b The block composition of alginate with G-blocks, M-blocks, and MG-blocks
2.2
Comonomer Sequence
2.2.1
Sequence Determination and Biological Distribution
The determination of the composition and sequential structure should be regarded as a mandatory task when dealing with alginate. This is because these two features are often key functional attributes which strongly affect the properties of the polysaccharide as well as its final performances both in biotechnological and food applications. It is certainly tempting to tackle the problem of the (co)monomer distribution in alginate by use of the typical statistical approach developed for synthetic (co)polymers. However, it has been ascertained that the monomer distribution along the alginate chain cannot be described by using Bernoullian statistics (Painter et al. 1968; Smidsrød and Whittington 1969). Moreover, a second-order Markov model can only provide an approximate description of the monomer sequence in the polysaccharide (Larsen et al. 1970). It is nowadays fully understood that the monomer distribution in alginates is the result of complex biochemical events, the correct identification and quantification of which require that comparatively long monomer sequences be known with extremely high accuracy. Certainly, the knowledge of the bare monomeric composition of alginate does not suffice to determine the sequential structure of alginate (Larsen et al. 1970). A complete characterization of the polysaccharide, from the point of view of the composition, can be efficiently achieved by high-resolution proton and
6
I. Donati and S. Paoletti
Fig. 2 1H-NMR spectrum (anomeric region) of an alginate isolated from Laminaria hyperborea (FG ~ 0.63). M-1M and M-1G represent the anomeric proton of an M residue neighboring another M residue or a G residue, respectively. MG-5M, GG-5M, and MG-5G refer to the H-5 proton of the central G residue in an MGM, GGM, or MGG triad, respectively. G-1 refers to the anomeric proton of G residues and GG-5G refers to the anomeric proton of G residues in G-blocks
carbon nuclear magnetic resonance (1H-NMR and 13C-NMR) (Grasdalen et al. 1977, 1979, 1981; Grasdalen 1983). An example of a typical 1H-NMR spectrum of alginate is reported in Fig. 2. Through this technique, the monomeric (FG, FM), diadic (nearest neighbors FGG, FGM, FMG, FMM), and G-centered triadic (next-nearest neighbors FGGG, F GGM, FMGG, F MGM) fractions can be calculated. In particular, for long chains in which the end effects can be neglected, the following relations hold: FG + FM = 1, FG = FGG + FGM = FGG + FMG , FM = FMM + FGM = FMM + FMG , FMG = FGM , FGG + FMM + FGM + FMG = 1, FG = FGGG + FGGM + FMGG + FMGM , FGGM = FMGG , FMG = FGM = FGGM + FMGM .
Material Properties of Alginates
7
From the relations above, the average length of G-blocks (NG>1) can be evaluated as N G >1 =
FG − FMGM . FGGM
It should be noted, however, that the 1H-NMR analysis provides an evaluation of the block chain length averaged over the whole G-block population. It follows that this technique, although very powerful, cannot provide information on the distribution of the G-block lengths within the polysaccharide structure. Very recently, a new approach based on the use of specific alginate-degrading enzymes (alginate lyases) coupled with a range of analytical techniques (electrospray ionization mass spectrometry, high-performance anion-exchange chromatography with pulsed amperometric detection, and matrix-assisted laser desorption/ionization time of flight mass spectrometry) allowed these limitations to be overcome by assessing the real length of the block structures in alginates (Holtan et al. 2006). The determination of the M-centered triads represents a more demanding issue since the 13C-NMR spectrum of the polysaccharides is required. The lack of this information hampers the exact determination of the average length of the MG-blocks in alginates. However, an estimation can be accomplished by considering the ratio FMGM/FGGM for the different polysaccharide samples (Grasdalen et al. 1981). A parallel and alternative characterization method exploiting circular dichroism (CD) has also been developed. It is based on the determination of the CD spectrum arising from each of the different sequential structures composing alginate, i.e., [q]GG(l) for the G-blocks, [q]MM(l) for the M-blocks, and [q]MG+GM(l) for the alternating sequences. Once these have been determined by using (nearly) pure M-blocks, G-blocks, and MG-blocks, the CD spectrum of an alginate sample, [q]alg(l), is considered as a linear combination of the three contributions above weighted for their fraction (Eq. 1) (Morris et al. 1980; Donati et al. 2003b). In such a way the diadic composition of alginate can be calculated using a nondestructive and very accessible technique, although it is less informative than NMR (Donati et al. 2003b).
[q ]alg (l ) = FGG [q ]GG (l )+ FMM [q ]MM((l ) + FGM+MG [q ]GM+MG (l )).
(1)
The composition and sequential structure of alginates from algal sources can vary over the range (0.2 < FG < 0.75). In general, high contents of guluronic acid are found for alginates from L. hyperborea plants, while alginates isolated from M. pyrifera, A. nodosum, and L. digitata are characterized by a lower content of G-blocks (Smidsrød and Skjåk-Bræk 1990; Draget et al. 2005). Alginates isolated from bacterial strains such as Azotobacter and Pseudomonas show more extreme and variable composition. While alginate obtained from the former bacterium is a real block copolymer, thus resembling the seaweed product,
8
I. Donati and S. Paoletti
alginates isolated from Pseudomonas species are characterized by the total lack of contiguous G residues. Even in alginates isolated from P. aeruginosa composed of almost 50% guluronic acid, the only G-containing sequence detected is the regular alternating structure (Sherbrock-Cox et al. 1984). This feature is also common to P. fluorescens and P. putida. An additional complication found in bacterial alginates is represented by the presence of O-acetyl groups, which invariably affect functional and material properties (ion binding and swelling) of the alginate itself (Skjåk-Bræk et al. 1986a; Ertesvåg et al. 1996). Finally, an alginate composed exclusively of mannuronic acid (mannuronan) can be isolated from an epimerase-negative mutant (AlgG−) of P. fluorescens (Gimmestad et al. 2003). An overview of the composition of the seaweed and bacterial alginates is reported in Table 1.
2.2.2
Modification of Composition and Sequence
Alginates, both in seaweed and bacteria, are produced by a postpolymerization reaction of mannuronan which involves an inversion of the configuration at the C-5 position of the uronic acid without breaking of the glycosidic bond. It is not within the scope of this chapter to describe in detail those enzymes. It suffices to stress that seven epimerases were found to be encoded in the genome of the alginate-producing bacterium A. vinelandii. These enzymes, named “AlgE1” to “AlgE7,” were sequenced, cloned, and overexpressed in Escherichia coli strains (Ertesvåg et al. 1994, 1996; Svanem et al. 1999). In such a way, after purification from the cells by sonication, centrifugation, filtration, and ion-exchange chromatography, large amounts of seven different epimerases were produced, allowing for compositional modifications of natural alginates. Each of these recombinant mannuronan C-5 epimerases generates specific nonrandom epimerization patterns, thus allowing the introduction of G residues into the alginate chains in different block structures. For example, the C-5 epimerase AlgE4 is able to convert homopolymeric M-blocks into strictly alternating copolymeric sequences, while the AlgE6 epimerase is a G-block-forming enzyme (Høidal et al. 1999; Hartmann et al. 2002a, b; Holtan et al. 2006). AlgE1 is characterized by the ability to introducing both MG-blocks and G-blocks into a mannuronan sample (Ertesvåg et al. 1998). The seven epimerases mentioned above also showed differences as to the substrate specificity. For example, AlgE5 has been shown to prefer MG-blocks as a substrate, while AlgE1 and AlgE6 prefer mannuronan or substrates containing G-blocks (Holtan et al. 2006). The in vitro use of such enzymes makes it possible to produce perfectly tailored alginate sequences not found in nature and opens up the possibility of controlling the length and distribution of blocks within the alginate chain, thus improving polysaccharide performances for several applications. In fact, it has been reported that mannuronan C-5 epimerases can lead to the
FG
FM
FGG
FMM
FGM,MG
FGGG
FMGM
FGGM,MGG NG>1
Algal Laminaria hyperborea (stipe) 0.63 0.37 0.52 0.26 0.11 0.48 0.07 0.05 15 Laminaria hyperborea (leaf) 0.49 0.51 0.31 0.32 0.19 0.25 0.13 0.05 8 Macrocystis pyrifera 0.42 0.58 0.20 0.37 0.21 0.16 0.17 0.04 6 Laminaria digitata 0.41 0.59 0.25 0.43 0.16 0.20 0.11 0.05 6 Lessonia nigrescens 0.41 0.59 0.22 0.40 0.19 0.17 0.14 0.05 6 Ascophyllum nodosum 0.41 0.59 0.22 0.38 0.21 0.13 0.14 0.07 5 Laminaria japonica 0.35 0.65 0.18 0.48 0.17 Durvillaea antarctica 0.32 0.68 0.16 0.51 0.17 0.11 0.12 0.05 4 Bacterial Azotobacter vinelandii 0.25–0.75 0.75–0.25 0.07–0.65 Pseudomonas aeruginosa 0 1 0 1 0 0 0 0 Pseudomonas fluorescens 0.40 0.60 0 0.2 0.4 0 0.4 0 FG denotes the fraction of alginate consisting of guluronic acid; FGG and FGGG indicate the fraction of alginate consisting of guluronic acid in blocks of dimers and trimers, respectively, whereas FMM indicates the fraction of alginate consisting of mannuronic diads; FGGM,MGG indicates the fraction of alginate which starts or ends with a block of guluronic acid; FGM,MG indicates the fraction of alginate consisting of mixed sequences of guluronic and mannuronic acid, with FMGM denoting the fraction of alginate consisting of two mannuronic acids interspaced with guluronic acid; NG>1 denotes the average length of guluronic acid blocks as NG>1= (FG−FMGM)/FGGM
Alginate source
Table 1 Chemical composition of some algal and bacterial alginates determined by high-resolution 1H-NMR spectroscopy
Material Properties of Alginates 9
10
I. Donati and S. Paoletti
production of alginates with enhanced gelling properties and stability (SkjåkBræk et al. 1986b; Draget et al. 2000). This enzymatic approach to composition tailoring was found to be extremely valuable in understanding the role of the different sequences in determining the overall alginate properties. In fact, by use of the epimerase technology, alginate samples composed of (almost) exclusively one block sequence were produced (Fig.3, Table 2) (Donati et al. 2005b). Recently, the combination of different epimerases allowed the design of engineered alginates where the block length and composition are finely tuned (Fig. 4, Table 2) (Mørch et al. 2007). In particular, knowledge of the mode of action and the specificity of the epimerases represents a fundamental tool for creating specific alginates with potentially any desired block composition. This novel and revolutionary approach can open up a completely new market for such a polysaccharide with outcomes covering advanced applications in the biomedical field.
2.3
Natural Substituents
O-Acetyl groups are invariably encountered in bacterial alginates and are the most characteristic feature distinguishing them from the algal ones. In alginates isolated from A. vinelandii and P. aeruginosa the acetyl groups are associated with the M residues and the ratio of the distribution between C-2 and C-3 is approximately 3:2, although some di-O-acetylated residues have been found (Table 3) (Skjåk-Bræk et al. 1986a). The role of O-acetyl groups in bacterial alginate is not completely understood. The presence of O-acetyl moieties has an inhibitory effect on mannuronan C-5 epimerization in A. vinelandii (Skjåk-Bræk et al. 1985) and on alginate lyase in some strains of P. aeruginosa (Linker and Evans 1984). It has been reported that acetyl substituents are necessary for the alginate isolated from P. aeruginosa to mediate cell aggregation into microcolonies in the early stages of biofilm development. In fact, an acetylation-defective mutant of such a bacterial strain was unable to form cell clusters and produced a thinner and more homogeneous biofilm (Tielen et al. 2005). The presence of O-acetyl groups in bacterial alginates brings about some notable differences in the solution properties of the polysaccharide and it is believed to perturb the stereoregular sequences, inducing a conformational change (Delben et al. 1982; Skjåk-Bræk et al. 1989b). Theoretical consideration yielded an estimated dimension increase of 20% after acetylation (Stokke et al. 1993) owing to a restriction of the conformational freedom of the polymer chains probably due to a hindered rotation of the monomers. From viscosity investigations, chain expansions were detected for degrees of acetylation of up to 11% (Skjåk-Bræk et al. 1989b).
O HO
OH
NaOOC
O HO
O
OH
O
HO
O
O
HO
NaOOC HO O
O
OH
O NaOOC HO
O HO
OH OH
O
OH
NaOOC
OH
OH O
O HO
NaOOC HO O O
Poly-M (Mannuronan)
O O
O
Poly-MG
NaOOC
O
O HO
OH
O
NaOOC
O
HO
NaOOC O
OH
OH
O
OH
O
NaOOC
OH NaOOC
OH O O
O HO
NaOOC
NaOOC
O
OH
O HO
OH
O
NaOOC
O
NaOOC
OH
O
OH
5.2
5.0
G-1
(ppm)
4.8
MG-5M
G-5G
4.6
M-1M
M-1G
4.4
Fig. 3 Structure and 1H-NMR (anomeric region) of polyG (guluronan), polyMG and polyM (mannuronan) (Table 2). See the legend to Fig. 2. (Adapted from Donati et al. 2005b)
HO
NaOOC HO O
NaOOC
HO
NaOOC
Poly-G (Guluronan)
12
I. Donati and S. Paoletti
Table 2 Chemical compositions of mannuronan and alginates obtained by means of AlgE epimerases Sample FG FM FGG FMM FGM,MG FGGG FMGM FGGM,MGG 0 PolyM (mannuronan)a 0.88 PolyG (guluronan)a PolyMG (polyalternating)a 0.46 0.48 PolyMG + AlgE1b 0.77 PolyMG + AlgE1b 0.55 PolyMG + AlgE6b 0.72 PolyMG + AlgE6b a Taken from Donati et al. (2005b) b Taken from Mørch et al. (2007)
2.4
1 0.12 0.54 0.52 0.23 0.45 0.28
0 0.82 0 0.05 0.59 0.18 0.50
1 0.06 0.08 0.10 0.05 0.08 0.05
0 0.06 0.46 0.43 0.18 0.37 0.22
0 0.78 0 0.04 0.56 0.15 0.47
0 0.03 0.46 0.41 0.16 0.35 0.21
0 0.04 0 0.01 0.03 0.03
0.03
Alginate as a Polyelectrolyte
Apart from the compositional issues characterizing alginate samples, a common feature of mannuronic acid and guluronic acid residues is that they both bear a carboxylic moiety that, at neutral pH, renders alginate a polyanion. The presence of fixed charges on the polysaccharide results in a notable influence on its overall dimension in solution, which is strongly influenced by the presence of added supporting salt. It is well known the ionic strength (I) of the solution is correlated to the Debye length (k−1), i.e., the distance at which the electrostatic interactions are screened out by the supporting salt, according to Eq. 2: k −1 ∝
1 I
.
(2)
The vast majority of the physicochemical analyses on alginates have been performed in the presence of 0.1 M sodium chloride. However, properties such as viscosity, the radius of gyration, and persistence length depend on the ionic strength of the medium used. In particular, the following linear relationship was proposed between the intrinsic viscosity of alginate, as polyanion, and the ionic strength (Fig. 5) (Fixman 1964; Smidsrød and Haug 1971):
[h ] = S
1 , I
(3)
with the slope of the curve, S, expressed as Eq. 4 (Fig. 5, inset) (Smidsrød and Haug 1971): u S = B [h ]0.1 , (4)
( )
where [h]0.1 is the intrinsic viscosity of the polyelectrolyte at 0.1 M sodium chloride, while B is a parameter inversely dependent on the rigidity of the polymer chain. n is
HO
HO O
G
NaOOC
M
NaOOC
M
HO O HO
OH
O
OH
G
NaOOC
O
G
O
O OH
OH
OH
O
OH
O HO
M
NaOOC
O
NaOOC
O
O
NaOOC
HO
NaOOC
HO
NaOOC O
OH
G
OH
O
M
HO O
M
G
G
HO
M
HO O
M
HO O O NaOOC
G
OH O
O
HO
O
M
HO
HO
M
HO O
NaOOC
NaOOC
OH
O
G
OH
O
HO
NaOOC
O
OH
NaOOC
O
M
HO O
AlgE1 /AlgE6
O
NaOOC
NaOOC
O OH
OH
HO
NaOOC O
AlgE4
M
O HO
O O HO OH
OH
NaOOC
O
HO
NaOOC
O
NaOOC
O
HO O
M
HO O
O
HO
O
NaOOC O
M
HO O O
Fig. 4 An enzymatic process to tailor-make alginates. Mannuronan is first converted into polyalternating alginate by the epimerase AlgE4. The use of AlgE1 or AlgE6 in a second reaction step leads to the introduction of G-blocks along the polysaccharide chain. (Adapted from Mørch et al. 2007)
HO
NaOOC
Material Properties of Alginates 13
14
I. Donati and S. Paoletti
Table 3 Composition and acetylation pattern of some bacterial alginates Source FG FM FGG FGM,MG FMM Total (%) C-2 (%) C-3 (%) C-2 and C-3 (%) A. vinelandii 0.67 0.33 0.54 0.13 P. aeruginosa 0 1 0 0 1 0 0 P. aeruginosa a 0
0.2 1 1
21 37 57
13 23 34
8 14 23
3 5
11
Data taken from Skjåk-Bræk et al. (1986) a After acetone reprecipitation
10
1
0.1
[η] (dL/g)
30
S (=Δ[η]/ΔI–0.5)
40
0.01 0.1
1
10
100
[η]0.1
20
10
0
5
10
15
I–0.5
Fig. 5 Dependence of the intrinsic viscosity [h] on the inverse of the square root of the ionic strength I for alginate. The dotted line represents the linear best fit of the experimental data. Inset: Dependence of the slope of the [h]-I−0.5 plot on the intrinsic viscosity measured in the presence of 0.1 M NaCl, [h]0.1. The dotted line represents the linear best fit of the experimental data. (Adapted from Smidsrød 1970; inset adapted from Smidsrød and Haug 1971)
usually assumed to be 1.3. Typically, B ranges from 0.04 to 0.05 in the case of alginates with high M content (Smidsrød and Haug 1971; Donati et al. 2004). The rigidity of the alginate chain is given by its total persistence length, qT , which is affected by the ionic strength used in the experimental determination. In the model of Odijk (Odijk and Houwaart 1978) qT of wormlike polyelectrolyte chains is the sum of two contributions: qT = q0 + qE .
(5)
In this equation, q0 is the intrinsic persistence length referring to the unperturbed dimensions of alginate chains. They are defined as the conditions in which
Material Properties of Alginates
15
excluded-volume effects and nonideal short-range interactions are counterbalanced. The second term, qE, is the contribution to the total persistence length due to the electrostatic interactions, which depend on the ionic strength; they are supposed to fade out at infinite ionic strength, neglecting any nonideal behavior of the supporting electrolyte. Typically, the intrinsic persistence length of alginate chains is around 12 nm, while the effect of the electrostatic interactions contributes, at an ionic strength of 0.17 M, an additional approximately 3 nm (Vold et al. 2006). The interactions of counterions with the polyanion (modeled as a homogeneous linear distribution of charges) are theoretically well described by the classical counterion condensation theory (Manning 1969a, b, 1977, 1978). It predicts that some of the monovalent cations will condense into a volume tightly surrounding the chain if the average distance between the charges falls below 7.135 Å (in water at 298 K): it corresponds to a value of the dimensionless linear charge density of 1 (for details, see Sect. 4.1). Looking at the molecular structure of alginate and considering lengths of 4.35 and 5.17 Å for guluronic acid and mannuronic acid repeating units, respectively, one can conclude that in both cases condensation of monovalent counterions is to be expected. This process conforms well to the theoretical prediction for an ideal system in which purely physical interactions hold. In contrast, when divalent ions are considered, an additional “nonbonding” chemical affinity of the condensed doubly charged ion for the polyelectrolyte (corresponding to a free energy contribution G0,aff) was shown to be present; they bring in speciesspecific effects, albeit without violating the general assumption of the condensation theory. In particular, theoretical calculations based on calorimetric measurement revealed that the affinity of nongelling divalent ions, such as Mg2+, decreased along the series G-block > MG-block > M-block, in parallel with the corresponding decreasing values of the linear charge density (Fig. 6) (Donati et al. 2006b). To some extent of analogy with the theory of “real” solutions with respect to “ideal” solutions, it is possible to address to such a “real” polyelectrolyte model as one in which chemical identity is taken into account in addition to the fundamental long-range nonspecific electrostatic interactions. Still, in the “real” polyelectrolyte the fundamental assumption of the counterion condensation theory is not violated: namely, that the fixed charge on the polyelectrolyte is preserved even in the presence of a fraction of condensed counterions.
2.5
Conformational Properties of the Alginate Chain
The influence of alginate composition on the flexibility of the polysaccharide chain has been discussed and analyzed quite extensively over the years. It was first reported by Smidsrød (1973) that the extension of the polysaccharide chain was dependent on its composition, with the intrinsic flexibility of the blocks decreasing in the order MG > MM > GG. This experimental result, based on light-scattering determinations, was also supported by different computer-modeling studies performed using a Monte Carlo method (Smidsrød and Whittington 1969; Stokke et al. 1993). Overall,
16
I. Donati and S. Paoletti 120
ΔH (cal•mol–1)
100 80 60 40 20 0 0.0
0.1
0.2
0.3 0.4 [Mg2+] / [Pol]ru
0.5
0.6
Fig. 6 Dependence on the ratio [Mg2+]/[Pol]ru of the experimental (symbols) and theoretical (lines) values of the enthalpy of mixing (DHmix) for polyG (guluronan, triangles and discontinuous lines), polyM (mannuronan, squares and continuous line) and polyMG (circles and continuous dots). The best fit between the experimental and theoretical curves was achieved by using an intrinsic free energy of affinity, G0,aff, of −3.55 kcal·mol−1 for polyG, −1.18 kcal·mol−1 for polyMG, and −0.30 kcal·mol−1 for polyM. (Adapted from Donati et al. 2006b)
such analyses, despite some variability and numerical difference in the predicted chain extension, all pointed to a flexibility of the block components of alginate chain that decreases in the order reported above. These results were supported by the experimental determination of the empirical B value for alginates with different compositions. In fact, the stiffness parameter B of alginate from M. pyrifera, which contains relevant amounts of MG-blocks and M-blocks, was higher than that isolated from L. hyperborea stipe, characterized by a higher frequency of G-blocks. In addition, both values were lower than that measured for a MG-block-rich alginate (Smidsrød et al. 1973). Since a higher value of the B parameter is associated with higher flexibility, the conclusion drawn was that G-blocks were intrinsically stiffer than the other sequences composing alginate (Smidsrød et al. 1973). Additional evidence on the effect of alginate composition on polymer chain extension came from the evaluation of the persistence length determined by light-scattering methods. In fact, alginate with different compositions displayed a decreasing persistence length upon increasing the M-block and MG-block fractions. In particular, average persistence lengths of 15 and 13 nm were found for alginate from L. hyperborea and A. nodosum, respectively, while for an MG-block-rich alginate, obtained from the acid-soluble fraction, a value as low as 9.5 nm was observed under the same conditions (Smidsrød et al. 1973). Gamini et al. (1992) compared two alginates isolated from M. pyrifera and L. hyperborea and found a value of the persistence length of 15 and 20 nm, respectively. Recently, molecular modeling investigations confirmed the differences in stiffness among the blocks present in alginate (Braccini et al. 1999).
Material Properties of Alginates
17
The overall scenario presented above has recently been revisited (Vold et al. 2006). The availability of pure alginate samples composed (almost) exclusively of one type of block sequence allowed reinspection of the intrinsic flexibility of the limiting components of the alginate chains. In particular, the empirical B parameter of an alginate composed exclusively of MG sequences was found to be markedly lower (B = 0.038) (Dentini et al. 2005) than that measured in the 1970s from acidsoluble alginate dominated by MG-blocks (B = 0.065). In parallel, the composition of engineered alginates, prepared by means of mannuronan C-5 epimerases, was shown to have a minimal effect on the B parameter, pointing to a limited effect of block sequences on the flexibility of the polysaccharide. This is also reflected in the evaluation of the persistence length of the different alginate samples containing largely one single block structure that was found to be around 15 nm independent of the composition (Vold et al. 2006). Reference to the conformational aspects of alginate in the specific case of its interaction with gelling divalent ions is given in Sect. 5.
3 3.1
Molecular Weight and Stability Molecular Weight and Molecular Weight Distribution of Native Alginates
Natural, bacterial, and enzymatically tailored alginate samples are polydisperse with respect to the molecular weight. This can be traced back to two main reasons: (1) the production of polysaccharides is not gene-encoded but it is under enzymatic control; (2) the extraction causes a substantial depolymerization. As a consequence, polydispersity in the degree of polymerization of these samples is more the rule than an exception. For polydisperse systems, the molecular weight is expressed by different averages over the whole distribution of the molecular — weights for the sample. The number-average molecular weight (M N) and the — weight-average molecular weight (M W) represent the most common averages used to address this point: MN =
∑NM ∑N i
i
i
MW
∑wM = ∑w i
i
i
i
i
i
,
(6)
i
∑NM = ∑NM i
i
2 i
i
i
i
,
(7)
where Ni is the number of molecules and wi is the weight of molecules having the given molecular weight Mi.
18
I. Donati and S. Paoletti
— Typically, commercial alginates have a weight-average molecular weight (M W) of approximately 200,000, but alginates with values as high as 400,000–500,000 are available. The molecular weight of native alginate is not known since degradation occurs during the extraction. — — The polydispersity index is given by the ratio M W/M N and provides an evaluation of the polydispersity of the sample. The polydispersity index is 1 in the case of a perfectly monodisperse polymer (proteins), while for polydisperse samples the — — relation M W > M N holds. Typically, alginate samples have a polydispersity index ranging from 1.5 to 3, although values as high as 6 have also been reported (Draget et al. 2005). Methods based on laser light scattering have proved their reliability in determining the molecular weight of alginate samples over a wide range of molecular sizes (Martinsen et al. 1991). Multiangle laser light scattering coupled with size-exclusion chromatography using a refractive index detector was found to be an efficient and reliable method for the determination of both the molecular weight averages and their distribution (Vold et al. 2006). In addition, the evaluation of the viscosity— average molecular weight (M V) can be performed by means of intrinsic viscosity determinations once the Mark–Houvink–Sakurada (MHS) parameters (K, a) are — — — known (see Eq. 12). For the latter average, it generally holds that M N < M V < M W, with MV = MW when a = 1.
3.2
Chain Degradation
The stability of an alginate molecule is strongly dependent on the conditions to which it is subjected, i.e., temperature, pH, and presence of contaminants. The glycosidic linkages between the sugar monomers of the polysaccharide are susceptible to cleavage both in acidic and in alkaline media. The first condition, which brings about a notable molecular weight reduction at pH values lower than 5, induces an acid hydrolysis of the glycosidic bond. In contrast, alkaline conditions cause chain scission by a b-elimination reaction at the glycosidic bond (Haug et al. 1963, 1967b). The rate of acid hydrolysis for alginate at pH around 3 or 4 was found to be much higher than that of neutral polysaccharides owing to intramolecular catalysis from the nondissociated carboxyl group which can form a favorable six-membered ring (Smidsrød et al. 1966). As to the alkaline conditions, it has been found that, in addition to hydroxyl ions, carbonate and phosphate ions function as catalysts in the b-elimination reaction owing to the general base-catalyzed nature of this reaction (Haug et al. 1967b). In both acid and alkaline conditions, the decrease in the degree of the polymerization follows first-order kinetics (valid for degrees of polymerization exceeding 10) (Haug et al. 1963): 1 1 − = k t, DPt DP0
(8)
Material Properties of Alginates
19
where DPt and DP0 are the degree of polymerization at time t and time 0, respectively. A very efficient way to characterize the degradation of alginate is to evaluate the time-dependent variation of the reciprocal of the intrinsic viscosity (see Sect. 4.2), i.e., 1/[h], which is proportional to the number of broken bonds in the polysaccharide chain. Such plots, which were found to be linear for [h] > 3, allow the overall stability of alginate in different pH conditions to be summarized (Fig. 7) (Haug et al. 1963). Alginate chains are very sensitive to the presence of free radicals that can be generated by the contaminants in the commercial samples (polyphenols) (Smidsrød et al. 1963a, b, 1967). Such radicals are known to depolymerize alginate by an oxidative–reductive depolymerization. The mechanism of degradation involves the auto-oxidation of reducing compounds with subsequent formation of peroxides (ROOH). These will then form hydroxyl radical species that could lower the degree of polymerization of the alginate chain at a very high rate (Smidsrød et al. 1963a, 1967). In addition, notable molecular weight reduction is encountered when alginate is treated with oxidants such as sodium periodate (Larsen and Painter 1969; Smidsrød and Painter 1973; Vold et al. 2006). In view of these considerations, safe handling conditions for alginate comprise a neutral pH (Haug and Larsen 1963) and limited heating as it is known that all the depolymerization reactions are boosted by temperature. It follows that sterilization using an autoclave should be avoided. Since the use of g-radiation could also be deleterious for the polysaccharide, filtration over 0.22-mm filters is the recommended sterilization procedure.
Δ(1 / [η])5h
0.1
0.01
1E-3 2
4
6
8 pH
10
12
14
Fig. 7 Degradation of alginate isolated from Laminaria digitata measured as the variation of the ratio 1/[h] after 5 h at different pH and at 68°C. The line is drawn to guide the eye. (Adapted from Haug et al. 1963)
20
I. Donati and S. Paoletti
In vivo biodegradation of alginates is performed by endolyases which catalyze the splitting of the 1–4 glycosidic bond via a b-elimination reaction. Modifications on the alginate chain, such as acetylation, lead to a net reduction of the enzyme activity. Endolyases are widely distributed in nature in prokaryotic and eukaryotic microorganisms and in bacteriophages that use alginate as a source of carbon. Lyases are found in the bacterial species producing alginate, such as A. vinelandii and P. aeruginosa, but not in the human gastrointestinal tract. Alginate-degrading enzymes have also been isolated from marine algae, including L. digitata. Like epimerases, endolyases exhibit sequence specificity in cleavage of either M or G residues in alginate chains (Gacesa 1992; Wong et al. 2000).
3.3
Chemical Modifications of Alginate
Several attempts to chemically modify alginate to vary its physicochemical properties have been reported. Among all these, the propylene glycol derivative of alginate (PGA) deserves special mention: it is an ester derivative obtained by reaction of propylene oxide with some uronic moieties of alginate. PGA is largely used as a thickener and stabilizer in several food applications mainly because of its solubility in acidic conditions and low sensitivity to the presence of divalent ions (typically Ca2+) (Yilmazer et al. 1991). In addition, the hydrophobic character introduced by the propylene glycol moiety allows this modified polysaccharide to function as a surfactant and emulsifier (Yilmazer and Kokini 1991). Recently, applications of PGA in tissue engineering have been proposed. Alginate is highly biocompatible and has very interesting features as to mass transport aspects; however, its biological inertness limits the applicability in the field of tissue engineering. In fact, the lack of bioadehesivity (Smetana 1993; Lee and Mooney 2001), i.e., the total absence of biological signals mimicking the extracellular matrix milieu and favoring the cell embedding, hampers the use of alginate hydrogels for the design of so-called third-generation biomaterials. The latter are able to directly intervene in cell recognition processes and therefore could bring about a notable benefit in tissue engineering. Several attempts have been made to overcome such a critical limitation. In particular, the chemical introduction of cell-specific signals, such as peptides and oligosaccharides, has been exploited (Rowley et al. 1999; Yang et al. 2002; Rowley and Mooney 2002; Donati et al. 2003a) and led to an enhancement of (favorable) interactions between the modified alginate and different cell types. However, the chemical modification of alginate brings about the drawback of reduced mechanical stability of its hydrogels. This can be safely traced back to the introduction of side-chain groups also (or even predominantly) on G residues (Donati et al. 2003a), which are the ones involved in gel formation. Recently, this issue was addressed and the use of a chemoenzymatic strategy allowed an engineered alginate modified exclusively on the non-gel-forming M residues to be obtained, and it hence had suitable mechanical properties for tissue engineering applications (Donati et al. 2005a).
Material Properties of Alginates
4 4.1
21
Alginate Solutions Thermodynamic Properties
Sodium alginate, like basically all the monovalent salts of alginate, is highly soluble in water owing to the favorable entropic contribution from the free (noncondensed) counterions. This consideration leads immediately to identification of the presence of a high amount of supporting salt as a limiting condition for the solubilization of the polyanion. In fact, the increase of the ionic strength of the alginate-containing solution causes the entropic gain arising from the polysaccharide counterions to level off. The addition of 1:1 salts, such as potassium chloride, will progressively disfavor the solubility of the alginate chain, leading eventually to phase separation. This salting-out effect can be efficiently used for the fractionation of alginates (Haug 1959a, b). In addition, the presence of ions could have a strong (negative) effect on the kinetics of alginate dissolution, since they act by reducing the water chemical potential difference between the alginate particle and the surrounding liquid. Concentrations of supporting salt even lower that 0.1 M slow down the kinetics of dissolution of alginate particles and limit their solubility. Besides this unspecific role of added salts contributing to the total ionic strength of the medium, the presence of specific divalent or multivalent ions can dramatically affect the solubility of alginate. This point will be discussed in more detail in the following section; here it suffices to say that divalent ions could induce a largescale chain aggregation mediating physical contacts between the polymer chains. This results in an extended network formation that leads to phase separation and, possibly, to hydrogel formation. The solubility of alginate may also be limited owing to the modification of the environmental conditions which do not imply variation in the ionic strength. In fact, the pH of the solution is a critical parameter controlling the solubility of the polysaccharide. It is known that the carboxyl groups of both mannuronic acid (pKa = 3.38) and guluronic acid (pKa = 3.65) can be protonated by addition of inorganic acids. Therefore, if the pH of the alginate-containing solution is lowered below the pKa of the constituting acids, phase separation or hydrogel formation occurs. Several investigations have focused on the pH dependence of alginic acid precipitation. It was found that molecular weight and composition play an important role in determining the pH of the precipitation of the polysaccharide (Haug and Larsen 1963; Haug et al. 1967a). In particular, MG-block rich alginates show a reduced tendency to phase-separate at acidic pH values when compared with G- and M-rich alginates, likely because of their higher degree of “conformational disorder” of the glycosidic bonds (Hartmann et al. 2006). The determination of the pKa values of uronic acids composing alginate can be traced back to the work of Haug, who also noticed that the pKa of alginate chains, (pKa)pol, does not differ significantly from that of its monomeric units. However, considering alginates isolated from different sources, their (pKa)pol value shows a
22
I. Donati and S. Paoletti
slight dependence on the polysaccharide composition. In addition, the (pKa)pol of alginate was found to depend slightly on the polymer concentration and on the ionic strength. In fact, the (pKa)pol for L. digitata alginate drops from 3.92 to 3.42 and that of L. hyperborea alginate drops from 4.41 to 3.74 on passing from a salt-free solution to aqueous 0.1 M sodium chloride (Haug 1964). Since alginate is an acidic polyelectrolyte, it is important to note that the (pKa)pol value, as well as the enthalpy of dissociation, ȈHdiss, depends on the degree of dissociation, a (Fig. 8): (pKa)pol(a) and ȈHdiss(a). Focusing on Fig. 8b, it is interesting to note that algal and bacterial (acetylated) alginates behave differently as to ȈHdiss(a) at low and high degrees of ionization. In particular, the ionized state of algal alginates is more favored by enthalpy than the one isolated from bacteria (Delben et al. 1982). Several calorimetric studies performed on both algal and bacterial alginates have focused on thermodynamic properties such as enthalpy of dilution, which provides a measure of the energy changes in the interaction of the solvated solute species upon dilution (Fig. 9). Owing to the polyelectrolytic nature of alginate, the enthalpy of dilution is a function of its linear charge density, x [x = e2/(DkBTb), where e is the value of the unitary charge and kB is Boltzmann’s constant, D is the bulk dielectric constant of the medium, T is the absolute temperature, and b is the (average) distance between fixed charges projected along the polyelectrolyte axis]. In particular, the (excess) electrostatic enthalpy change upon mixing the polymeric salt with pure solvent to change its concentration (Ce) from Ce,i to Ce,f, per mole of uronic acid group, is given by Eq. 9 (Paoletti et al. 1985): ⎛ k b k b ⎞ 1 1 − e −k f b ⎛ d ln D ⎞ ΔH dil = − RT ⎜1 + 1 − r x 2ln + k f bf + k i bi ⎟ , ( ) ⎟ −k i b ⎜ ⎝ 2 d ln T ⎠ 1− e e − 1 e − 1⎠ ⎝
a 5.0
ΔHdiss (kcal / molH+)
b
(pKa)pol
4.5 4.0 3.5 3.0
0.0
0.2
0.4
0.6 α
0.8
1.0
(9)
1 0 –1 –2 –3 0.0
0.2
0.4
0.6
0.8
1.0
α
Fig. 8 a Apparent dependence of (pKa)pol on the degree of ionization, a, for (squares) algal alginic acid titolated with NaOH in water at 25°C. b Enthalpy changes on proton dissociation (ȈHdiss) for acetylated (circles, degree of acetylation approximately 8%) and deacetylated (squares) alginic acid as a function of the degree of ionization, a, in water at 25°C. The lines are drawn to guide the eye. (Adapted from Delben et al. 1982)
Material Properties of Alginates
23
ΔHdil (cal•equiv–1)
0
–100 PolyG –200
–300 1.5
PolyM
2.0
2.5
3.0
3.5
4.0
–logCe
Fig. 9 Dependence of the heats of dilution, ȈHdil, on the negative logarithm of the final polymer concentration, Ce, for acetylated (open squares) and deacetylated (filled squares) sodium alginates (a ~ 1) in water at 25°C. The dotted lines are drawn to guide the eye. The heat of dilution, calculated according to Eq. 9 derived from the counterion condensation theory, for polyG (guluronan) and polyM (mannuronan) is also reported. (Adapted and modified from Paoletti et al. 1985)
where ki and kf are the reciprocal of the Debye length at initial and final ionic strength, respectively, and r is the fraction of monovalent ion condensed on the −1 polyelectrolyte (r = 1 − x ). Equation 9, derived from the counterion condensation theory, was found to hold in the case of algal alginate and to parallel with a very high accuracy its exothermic experimental enthalpy of dilution (Benegas et al. 1992). In particular, the slope of the ȈHdil versus –logCe experimental data points out that the algal copolymer falls in-between the theoretical predictions for the two limiting components of homopolymeric sodium mannuronate and guluronate (Fig. 9) (Paoletti et al. 1985). In contrast, when a bacterial (acetylated) alginate is considered, significant differences are detected. In particular, in the latter case the ȈHdil versus –logCe curve departs from the expected behavior in a way which is suggestive of a concomitant conformational change of the polymer. Additionally, for algal alginate samples, a theoretical approach based on the counterion condensation theory was found to correlate very well with the experimental enthalpy of mixing of alginate with monovalent ions, namely, Na+ (Donati et al. 2006b).
4.2
Nonequilibrium Properties
The viscosity of an alginate solution is a measure of the hydrodynamic volume of the polymer chains. As such, it is strongly affected by the average length of the chains in the ensemble (i.e., it depends on some average of the molecular weight)
24
I. Donati and S. Paoletti
and the conditions used (solvent, temperature, ionic strength), which collectively influence both extension (expansion) and flexibility of the polysaccharide sample. An increase of alginate molecular weight will obviously increase the viscosity of the alginate solution, providing some beneficial effects for several applications, in particular in the food industry and textile printing. However, the use of a high molecular weight alginate results in a solution which is difficult to handle (sterile filtration could become a major issue for such systems) and that could reduce cell viability (Kong et al. 2003), owing to the high shear forces required to mix it with cells for cell therapy. The intrinsic viscosity [h] of alginate is experimentally determined in dilute polymer solutions by means of the Huggins (Eq. 10) and Kraemer (Eq. 11) relations: hsp c
= [h ]+ k ′ [h ] c, 2
ln hrel 2 = [h ]− k ′′ [h ] c, c
(10)
(11)
where hsp, hrel, and c are the specific viscosity, the relativity viscosity, and the alginate concentration, respectively, while k¢ and k¢¢ are the Huggins and Kraemer constants, respectively. The value [h] is used to evaluate the molecular weight of the alginate sample through the MHS equation:
[h ] = KM wa .
(12)
Generally alginate is considered a semiflexible wormlike chain for which the a parameter of the MHS equation exceeds the theoretical value of a random-coil chain in q conditions (i.e., 0.5). Several authors have challenged the determination of the semiempirical K and a parameters for alginate (mainly for samples isolated from L. hyperborea) over a wide range of molecular weights. The MHS parameters were first determined as a function of the ionic strength by Smidsrød (1970) (Table 4). It is important to note that the a parameter, describing the polymer chain extension and conformation, exceeds the value expected for Gaussian coils in the q condition even for the extrapolation at infinite supporting salt concentration (I→∞). This effect has been considered as a consequence of the intrinsic rigidity of the polysaccharide chain. Values of a above 0.5 are usually attributed to excluded-volume effects, resulting in an expansion factor over the unperturbed dimensions which increases with molecular weight. The data obtained by Smidsrød were basically confirmed by Mackie et al. (1980), who proposed a set of different MHS semiempirical parameters depending on the ionic strength used in the analysis and on the composition of the alginate samples. Additional investigations reported a value for a in the MHS equation ranging from 0.92 for a M-rich alginate to 1.13 for a G-rich alginate (Martinsen et al. 1991). In the cases reported, viscosity measurements were performed in combination with
Material Properties of Alginates
25
Table 4 Semiempirical parameter a obtained for alginate Smidsrød (1970) Ionic strength L. digitata 0.01 1.15 0.1 1 0.2 0.5 1 0.87 ¥ 0.84
Mackie et al. (1980) L. hyper- A. vineborea landii 1.31 1.08 1.11 0.91 1.06 0.86 1.00 0.82 0.80
Berth (1992)
0.90
Martinsen et al. (1991) L. hyperbo- M. pyrifrea era
Vold et al. (2006) L. hyperborea (stipe or leaf)
1.13
1.01
0.92
0.73
wide-angle laser light scattering and low-angle laser light scattering analyses to determine the molecular weight of the samples. Table 4 summarizes the different values of the semiempirical parameter a reported for different alginate samples. Recently, Vold et al. (2006) reported the following relation for alginate samples to be valid for practical reasons for all the alginate samples regardless of the source:
[h ] = 0.00504 M w1.01 (0.1M NaCl , T = 20° C).
(13)
The use of a size-exclusion chromatography with online multiangle laser light scattering and viscometry also suggested that, in the case of L. hyperborea alginate, two separate regions in the linear log[h]-logMw relationship coexist depending on the alginate molecular weight range (Eqs. 14, 15) (Vold et al. 2007). 20, 000 < M w < 100, 000 : [h ] = 0.0051M w1.00
(14)
100,000 < M w < 300,000: [h ] = 0.0349 M w0.83 .
(15)
The use of light-scattering techniques allowed the determination of the relationship between the radius of gyration of alginate and its molecular weight as Rg2
1/ 2
= 0.095M w0.54 ( M w > 65, 000) (Smidsrød 1970)
(16)
or as Rg2
1/ 2
= 0.0352 M w0.60 . (Vold et al. 2006)
(17)
The linearity of the relationship between the reduced specific viscosity and alginate concentration (Eq. 10) is generally fulfilled until a critical concentration, C*, is reached, which corresponds to a coil overlap factor, f = [h] CAlg, of 1. Once the semidilute regime has been reached, the relative viscosity scales strongly with the polymer concentration, namely, according to hspaCA3.3lg− 3.1 (Morris et al. 1981;
26
I. Donati and S. Paoletti
Donati et al. 2007). Alginate solutions in the semidilute regime show, additionally, a marked shear-thinning effect which leads to a reduction of the viscosity upon increase of the shear rate applied. As a consequence, the (viscosity) relaxation time of alginate solution scales as t relax a CA1.8lg , thus substantially in good agreement with the value expected for a semidilute entangled system, i.e., trelax μ C3/2Alg.
5
Ion-Binding Properties
The most relevant feature of alginate, from both the industrial and the biotechnological point of view, is connected with its ability to efficiently bind divalent cations, such as Ca2+, Sr2+, and Ba2+ to name a few, leading eventually to hydrogel formation. This polysaccharide has shown a highly selective binding ability and the affinity towards different alkaline-earth divalent ions was found to increase in the order Mg2+<<Mn2+
Sr > Ca >> Mg MM-blocks: Ba > Sr » Ca » Mg MG-blocks: Ba » Sr » Ca » Mg
Material Properties of Alginates
27
100
KCaMg
1
10 2 3 1 0.0
0.2
0.6
0.4
0.8
1.0
XCa
Fig. 10 Selectivity coefficients, KMgCa, as a function of ionic composition (XCa) for different alginate fragments. Open circles dialysis of the fragments in their Na+ form. Filled circles dialysis first against 0.2 M CaCl2, then against mixtures of CaCl2 and MgCl2. Curve 1 fragments with 90% guluronate residues. Curve 2 alternating fragment with 38% guluronate residues. Curve 3 fragment with 90% mannuronate residues. The lines are drawn to guide the eye. (Adapted from Smidsrød 1974)
Overall, each cross-linking ion (i.e., Ca2+) interacts with two adjacent G residues as well as with two G residues in the opposing chain, thus inducing the formation of junction zones. The latter represent the physical interactions connecting alginate chains together and leading to the formation of the hydrogel. Therefore, the junction zones in the gel can be envisaged as “nanocrystallites” flooded by the “sticky sea” of elastically active chains. On the basis of dialysis experiments, which showed a 4:1 ratio between G residues and calcium ions, a schematic sketch of the possible junction formed in the G-rich alginate has been proposed (Fig. 11). The formation of the junction, and eventually of the hydrogel, upon addition of the divalent ion can be described as a cooperative process with an unfavorable binding of the first ion and a more favorable binding of the following ones (zipper mechanism). It has been reported that in the case of Ca2+, a minimum of eight to 20 adjacent G units are required for the formation of a stable junction. Also in this case a strict parallel with pectate holds (Kohn and Luknar 1977; Donati et al. 2006a). However, the minimum length of the G-block required for the binding of a cross-linking ion decreases upon increasing the affinity of the latter. Direct verification of the structural features of the egg-box model is strongly hampered by the lack of crystalline and well-resolved X-ray structures of alginate in the presence of divalent ions such as Ca2+ or Sr2+. Most of the crystallographic data
28
I. Donati and S. Paoletti
a
coo– HO
coo–
HO HO
O
HO
O
O
O
O
O
HO
coo–
HO –
O
O
HO
HO
Ca2+
ooc
ooc
OH OH
coo
Ca2+
–
OH OH
O
O O
O
O
–
O O
OH –
OH
O
ooc
OH
OH –
ooc
b
Ca2+
Fig. 11 Binding of divalent cations by alginate – the “egg-box” model. a Chelation of divalent cations. b Interchain junction formation
available arise from the acidic form of the polysaccharide in the fiber form: the derived structural features are used to predict the crystal structure and ion coordination geometry of Ca2+–alginate junctions in the physical gels (Atkins et al. 1973a, b). Over the years, refinements of the egg-box model have been attempted by means of a reinspection of the original X-ray diffraction data and a slightly modified crystal structure for the calcium-induced junction was proposed on the basis of the “formal” replacement of one water molecule with the divalent ion found in the unit cell of the free acid form of alginate (Arnott et al. 2000). A good-resolution X-ray data for Ca2+–alginate junctions has, however, been achieved recently. These novel experimental data recognized the 2/1 helical structure of the G-blocks in the junctions and helped determine both the stoichiometry and the structural features of the complex between G-blocks and calcium (Sikorski et al. 2007). Additionally, the description of the calcium-induced interchain association of the G-blocks in the hydrogel junctions has been attempted by means of molecular modeling. In this case, the positioning and the coordination of the ion differed, to some extent, from the those in original egg-box (Braccini and Pérez 2001) model and these conclusions were supported by NMR spectroscopy (Steginsky et al. 1992). From the inspection of the data reported above, it should be concluded that the basic features of the ion-induced alginate junctions were well represented in the original formulation of the egg-box model provided by Morris, Rees and Smidsrød.
Material Properties of Alginates
29
Therefore, the description of alginate junctions as composed of dimers or of few laterally associated polymer strands still holds after 30 years. This overall picture has been reconsidered and rediscussed in view of smallangle X-ray scattering experiments that suggested a lateral association of calciumcontaining junctions, far beyond a pure dimerization, takes place once an excess of ion has been provided (Stokke et al. 2000). However, recent X-ray analyses have indicated that alginate junctions are largely composed of polysaccharide chain dimers held together by Ca2+ coordination in close resemblance to the egg-box model, with some limited lateral association occurring through unspecific electrostatic interactions (Sikorski et al. 2007). The divalent ion binding by alginate chains has also been screened by calorimetric and dilatometric methods. In particular, the enthalpy of mixing of alginate with divalent cations of high affinity, such as Ca2+ and Pb2+, was exothermic and characterized by a sigmoid shape (Fig. 12) (Cesàro et al. 1988b). This result revealed a contribution of conformational origin (disorder-to-order transition) deriving from the G-rich component of the polysaccharide chain, as corroborated by CD data (Morris et al. 1975; Cesàro et al. 1988a). In addition, such results pointed to the cooperativity of the ion-induced junction formation. Moreover, dilatometric studies revealed that alginate-ion interactions are associated with a positive volume change. It stems from a release of a large number of water molecules (desolvation) from the
0.5
24 22 20
0.0
16 14 12 10
–0.5
8
Δ v (cm3 • mol–1)
ΔH (kcal•mol–1)
18
6 4 2 –1.0 0.0
0.2
0.4
0.6
0.8
0 1.0
RJ ([M2+] / [Pol]ru)
Fig. 12 Enthalpy (left y-scale, filled symbols) and volume (right y-scale, open symbols) changes upon mixing of alginate with Pb2+ (squares) and Ca2+ (triangles) in aqueous NaClO4 (0.05 M) as a function of Rj (the ratio of the number of moles of ion per monomole of polymer). The lines are drawn to guide the eye. (Adapted from Cesàro et al. 1988b)
30
I. Donati and S. Paoletti
solvation spheres of both the divalent ion and the chelating polymer chain, bringing about a favorable increase in entropy (Fig. 12). The combination of the quoted results clearly shows that the extremely high affinity of alginate for ions such as Ca2+ or Pb2+ must be traced back to both negative enthalpy changes and positive entropy changes. A thermodynamic explanation is then given for the only slight temperature dependence of the stability of lyotropic calcium alginate gels, at variance with, for example, the thermolyotropic gels of the sulfated polysaccharides k-carrageenan and i-carrageenan with K+. The availability of alginate epimerases, which are able to produce alginate polymers composed (almost) exclusively of one block structure represents a very powerful tool to underline the role of the different sequences with respect to their ion-binding properties. In particular, the epimerization of mannuronan with C-5 epimerase AlgE4 allows the production of a pure polyalternating alginate, thus composed exclusively of MG-blocks, which, in spite of the complete lack of G-blocks, was able to form stable hydrogels upon addition of calcium. Moreover, the elongation of alternating sequences in natural alginates, obtained by means of the AlgE4 epimerase, was found to induce a marked increase in the mechanical properties and syneresis of the calcium hydrogels. These novel findings allowed the proposal of the existence, in alginate gels, of three possible junctions. More specifically, the direct involvement of repeating MG units in the formation of both mixed MG/GG and pure MG/MG junctions has been suggested (Fig. 13) (Donati et al. 2005b). Thanks to the availability of alginates with tailor-made block compositions, the selectivity of binding by the different block structures has been revised, revealing the binding ability of M-blocks towards Ba2+ ions and of MG-blocks towards Ca2+ and Sr2+ ions (Mørch et al. 2006). The overall picture resulting from these analyses helps shed some light on the different effects of the same ion on natural alginates characterized by a different composition (Mørch et al. 2006). These studies have a profound impact in the alginate field as it is now clear that detailed knowledge of the composition and the ion-binding properties is fundamental for the selection of the alginate in order to have the desired material properties.
6
Alginate Gels
6.1
Gel Formation Properties
6.1.1
Acid Gels
Alginate solutions treated at a pH below the pKa of uronic acid can, under certain conditions, lead to the formation of acid gels. The latter, in contrast to ionic gels, are equilibrium gels, as proved by swelling studies, and are stabilized by hydrogen bonding (Draget et al. 1994; Smidsrød and Draget 1996). It has been shown that the gel strength of acid gels becomes independent at a pH below of 2.5. However, acid hydrogels
Material Properties of Alginates
31
Fig. 13 The three possible junctions in alginate hydrogels: a GG/GG junction, b MG/MG junction, and c GG/MG junction
resemble ionic ones in the sense that the gel strength seems to be correlated with the content of G-blocks in the alginate chain, pointing to a possible cooperativity in the gel formation process. With the exception of some pharmaceutical uses, the number of applications for alginic gels is so far limited (Draget et al. 2005).
6.1.2
Ionic Gels
The treatment of an alginate semidilute solution (1–2%) with a proper divalent cation, i.e., Ca2+, results in the formation of a ionotropic (lyotropic) gel. The ion-binding ability of alginate is the basic feature controlling hydrogel formation. Overall, an alginate gel can be depicted as a continuous network swollen with water where
32
I. Donati and S. Paoletti
physical cross-links, represented by the ion-induced junction zones, hold together different polysaccharide chains. The formation of the hydrogel is favored by the fact that Ca2+–alginate gels have a markedly higher affinity than Na+–alginate solution towards Ca2+ ions. This has been explained theoretically as a near-neighbor autocooperative process which predicts that affinity towards a specific ion increases with increasing content of the same ion in the medium (Smidsrød 1974). The gelling kinetics is strongly dependent on the method used for the introduction of the cross-linking ion and hence homogeneous or inhomogeneous hydrogel can be obtained (Fig. 14).
6.1.2.1
Homogeneous Alginate Hydrogels
To achieve a homogeneous distribution of alginate within the hydrogel, a controlled introduction or release of the cross-linking ion has to be accomplished. Over the years, the internal gelation method has been established as a reliable procedure for achieving a release of calcium ions in a controlled fashion (Draget et al. 1991). This is based on the use of an inactivated form of calcium (CaCO3, Ca-EDTA, calcium citrate) that is mixed with the alginate solution. The controlled release of the crosslinking ion is usually obtained by means of a decrease of pH caused by the addition of a slowly hydrolyzing molecule such as d-gluconolactone (GDL). The use of calcium carbonate and GDL, once a molar ratio of GDL to CaCO3 of 2 is maintained, guarantees only slightly acidic pH in the final hydrogel, thus with a negligible contribution from acidic alginate gels. In contrast, when Ca-EDTA is used, the
Hydrogel boundary
Calcium Reservoir
Alginate
CaCO3
H2CO3
2H+
GDL
Diffusion setting
Internal gelation
Fig. 14 The diffusion setting and internal gelation methods
H2O + CO2
Material Properties of Alginates
33
presence of some acidic gel cannot be ruled out. Generally, the internally set gels are prepared in calcium-limited conditions to favor handling and the preparation of homogeneous gels. It follows that their mechanical performances will be lower when compared with those of saturated calcium alginate gels. The kinetics of the gelation process has been studied extensively and it has been found to be influenced by alginate concentration, composition, and particle size of the calcium salt (i.e., CaCO3) (Draget et al. 1991). The influence of alternating sequences in the gelling kinetics on the homogeneous alginate networks has also been explored. In particular, it has been reported that the longer the MG-blocks are in the alginate chains, the faster is the gel setting (Draget et al. 2000).
6.1.2.2
Inhomogeneous Alginate Hydrogels
Hydrogels characterized by an inhomogeneous distribution of the polysaccharide can be obtained by means of the diffusion setting. This is performed by allowing a cross-linking ion (e.g., Ca2+) to diffuse from a larger outer reservoir into an alginate solution. In this case, very rapid gelling kinetics occurs at the front of the diffusing ion, resulting in an instantaneous formation of the hydrogel. This approach has been largely used in the preparation of alginate beads for the entrapment of biological material and for cell immobilization (Strand et al. 2002). In the case of the setting of gel beads obtained from diffusion, highly inhomogeneous alginate distributions have been reported, with a fivefold increase in alginate concentration at the surface and a very low concentration (virtually zero) in the center of the gel. Hence, the polymer is mainly located on the external part of the bead, while a very soft and liquid core is found in its central part. Inhomogeneity of alginate beads has been studied by several methods, including synchrotron radiation induced X-ray emission, magnetic resonance imaging, confocal microscopy, and Raman spectroscopy (Thu et al. 2000; Strand et al. 2003; Heinemann et al. 2005). The polymer gradient within the gel can be controlled by careful selection of alginate molecular weight, concentration, and composition, as well as of the cross-linking ion used and of its concentration (Thu et al. 2000; Strand et al. 2003; Mørch et al. 2006). In general, low molecular weight alginate and (relatively) low concentrations of the gelling ion will give the highest polysaccharide gradient throughout the hydrogel. In contrast, a high molecular weight alginate and a high concentration of gelling ions will increase the homogeneity of the hydrogel (Skjåk-Bræk et al. 1989a). In general, a high inhomogeneity of the hydrogel may be preferred in the preparation of microcapsules characterized by lower porosity (Martinsen et al. 1992) and higher resistance against swelling (Thu et al. 1996b). From the theoretical point of view, the strong and virtually irreversible site binding of the cross-linking ion by two facing G-blocks in the alginate chain has been regarded as the main cause of the uneven distribution of alginate within the gel. The cross-linking ions rapidly diffuse towards the center of the droplet owing to the difference in the concentration between the reservoir and the alginate bead. The diffusion of gelling ions will create a sharp gelling zone that moves from the surface towards
34
I. Donati and S. Paoletti
the center of the gel. The activity of alginate (and of the gelling ion) will be zero in this zone, and alginate molecules will diffuse from the internal, nongelled part of the gelling body towards the zero-activity region. However, the diffusion of alginate (a polyanion) is enhanced, with respect to self-diffusion on a noncharged macromolecule, by the coupled diffusion of the counterion (Na+). The latter, owing to the unbalanced concentration between the alginate bead and the reservoir, will rapidly diffuse towards the gelling front, dragging the alginate molecule for electroneutrality reasons. The addition of a nongelling ion, such as sodium chloride, to the reservoir (gelling bath) containing the cross-linking ion reduces the diffusion of the counterion of alginate towards the external part of the bead. As a consequence, the coupled diffusion between the counterions and the macromolecule is hampered and the resulting gel is homogeneous with a uniform distribution of the polysaccharide throughout the bead (Skjåk-Bræk et al. 1989a; Strand et al. 2003).
6.1.3
Syneresis
“Syneresis” is the term used to describe the slow shrinking of alginate upon increasing the divalent cross-linking ion concentration. Macroscopically, syneresis is detected as a release of water from the gel with a consequent decrease in its dimensions and increase in polymer concentration (Fig. 15). This partial collapse of the network increases the stability of the hydrogel and reduces its porosity: both features are highly advantageous in the case of the use of alginate for microencapsulation for cell transplantation (Martinsen et al. 1992; Thu et al. 1996b). The alginate gel shrinkage has been reported to depend on the composition of the alginate sample. In fact, alginates containing long G-blocks will shrink less than alginates with shorter G-blocks owing to the formation of strong irreversible junctions, thus hindering reorganization of the network structure (Martinsen et al. 1989). In addition, syneresis increases with increasing calcium concentration, gelling time (Martinsen et al. 1989), and alginate molecular weight (Draget et al. 2001). Although macroscopically very well characterized, the molecular details of syneresis are still debated. Undoubtedly, MG-blocks in alginate chain do play an important role in determining gel shrinkage. In fact, a correlation between the frequency of alternating sequences in the polysaccharide and the loss of water by the hydrogel has been reported. This was traced back to the higher flexibility of the MG-blocks, with respect to the other blocks present in alginate, thus allowing a denser packing of the gel network (Draget et al. 2001). As a consequence, lateral association of calcium-junction zones, beyond dimerization, has been suggested to account for the syneresis of alginate hydrogels (Stokke et al. 2000). More recently, it has been underlined that the length of the MG-blocks, rather than their overall amount, determines the extent of the collapse of the hydrogel. Therefore, syneresis was thought to be a result of a “zipping” of long alternating sequences in secondary MG/MG junctions upon increasing the Ca2+ concentration (Donati et al. 2005b).
Material Properties of Alginates
35
7
Ca2+−Alg in beads (%W/V)
6 5 4 3 2 1 0
0
1
2
3
4
5
Na+−Alg in solution (%W/V)
Fig. 15 Calium alginate concentration in the gel bead versus sodium alginate concentration in the solution for samples isolated from Macrocystis pyrifera (filled squares) and L. hyperborea stipe (open squares). (Adapted from Martinsen et al. 1989)
It is likely that both mechanisms contribute, to a different extent, to the final shrinkage of the alginate hydrogel.
6.2
Hydrogel Stability
Alginate hydrogels are easily obtained by using a proper cross-linking ion, which, being chelated by the G sequences in the alginate chain, induces the formation of the three-dimensional network. As such, the presence of a semisolid state is guaranteed by the physical interactions between the ion and some of the uronic moieties composing the polysaccharide chain. These interactions are by far less resistant than chemical linkages and their depletion markedly affects the hydrogel stability. It follows that the major limitation to the use of alginate gel is represented by the presence in the medium of compounds able to chelate divalent ions as well as by a high concentration of competing ions. Treatment of alginate gels with one of these two systems causes gel swelling, variation of its porosity, and eventually dissolution. It is well known that the volume of a polyelectrolyte hydrogel in a solvent is determined by the equilibrium between the osmotic pressure (P1osm), which tends to increase the overall dimensions of the hydrogel, and the elastic retractive force (P1el), which, in contrast, acts to impede volume expansions: Π1osm + Π1el = 0.
(18)
36
I. Donati and S. Paoletti
From the theoretical point of view, the same concept can be rephrased by stating that, under equilibrium conditions, the chemical potential of the solvent inside the gel (msolv,In) equals that outside the gel (msolv,Out), i.e., in the surrounding solution. For a polyelectrolyte gel this reads msolv,In = msolv,Out → Δmsolv = Δm mix + Δmion + Δm el .
(19)
The polymer–solvent mixing term, Dmmix, basically depends on the interactions between the polymer and the solvent (hydration of the polymer) (Flory 1953; Hoffman 2002). This term has been reported to be negligible in the case of alginate gels (Moe et al. 1993). An osmotic contribution facilitating gel swelling arises from the difference in chemical potential of the water owing to uneven distribution of ions between the inside and outside of the gel, Dmion. In this case, the osmotic swelling force results from the presence of the counterions of the fixed charges on alginate. These mobile counterions cannot leave the hydrogel, because the electroneutrality condition must always be fulfilled, and therefore their concentration will always be greater in the gel than outside. Hence, an osmotic pressure is exerted on the semipermeable membrane of the hydrogel, causing an increase of its volume. The swelling of the hydrogel is opposed by the physical cross-links (network junctions) which cause an elastic network retraction force, Dmel (Moe et al. 1993; Hoffman 2002). The overall equilibrium volume of the polyelectrolyte hydrogel is determined by the balance of the contributions reported. In view of these considerations, it can be easily understood that the presence of competing ions, as well as of chelating agents, acts in two ways, both contributing to hydrogel network swelling. First, calcium ions are removed from the junctions, hence decreasing the number of physical cross-links and reducing the elastic network retraction force. Second, the replacement of the divalent cross-linking ions with monovalent ones increases the number of dissociable counterions (one calcium ion is replaced by two sodium ions) and thus the osmotic pressure. From what was reported above, it is clear that phosphate, citrate, lactate, and EDTA should be avoided when dealing with calcium alginate hydrogels since the presence of such chelating agents will remove the cross-linking ion from the network junctions. Similarly, high concentrations of competing ions, such as Na+ in physiological saline solution (i.e., 0.15 M NaCl), will hamper the stability of the network. Several strategies have been devised to overcome the limited stability of alginate gels in physiological conditions. As a first choice, the replacement of calcium ions with stronger binding ones, such as Ba2+ or Sr2+, coupled with the use of alginate chains containing a high fraction of G residues has been reported (Smidsrød et al. 1972; Mørch et al. 2006). A notable increase in the stability of the alginate hydrogel was noticed and, although the toxicity of these ions used for biomedical applications still represents a concern, no Ba2+ leakage was detected from hydrogels prepared from high-G alginate after extensive rinsing (Thu et al. 1996a). In addition to this approach, the use of polycations, such as poly(l-lysine), poly(l-ornithine),
Material Properties of Alginates
37
chitosan, and polyethyleneimine, for the preparation of a polyanion-polycation complex membrane has been explored (Lim and Sun 1980; Tanaka et al. 1984a; Thu et al. 1996a; Gåserød et al. 1998; Taqieddin and Amiji 2004; de Vos et al. 2006; Blasi et al. 2006; Lawrie et al. 2007). The strong electrostatic binding taking place between the two oppositely charged polyelectrolytes account for an increase of the physical interactions and for a partial discharging of the polymer. Very stable beads can be prepared depending of the alginate hydrogel dimensions, the concentration of the polycation, and the exposure time. However, the presence of a polycation on the surface of the hydrogel, when implanted as a bioreactor in the body, might evoke a strong immunitary reaction with the formation of a fibrotic capsule around the bead which impedes the correct exchange of nutrients and oxygen to the biological material entrapped within the network. Although this adverse effect can be reduced by an additional coating with a polyanion, a complete suppression of the inflammatory response by the polycation is still an open issue. Another approach for stabilizing alginate gels comprises the introduction of covalent cross-links in addition to the physical (ion-induced) ones. Various chemical techniques have been attempted, including covalent grafting of alginate with synthetic polymers, a combination of covalent and ionic cross-links on the polysaccharide, as well as the direct reticulation of poly(l-lysine) on alginate (Hertzberg et al. 1995; Dusseault et al. 2005, 2008). A drawback in the methods based on the direct chemical modification of alginate is the lack of selectivity. In fact, grafting of side chain groups occurs mainly on the G residues of the polysaccharide, thus hampering gel formation by the modified alginate. A new approach has recently been shown to be able to overcome this limit by using a combination of chemical grafting of methacrylate moieties onto mannuronan, composed exclusively of M-blocks, and of different mannuronan C-5 epimerases (Rokstad et al. 2006). Finally, the composition of the alginate chain was found to have a profound impact on dimensional stability. In particular, the presence of long alternating sequences in the polymer, introduced by an enzymatic approach, leads to very stable alginate gel beads that do not swell upon treatment with saline solution (Fig. 16) (Mørch et al. 2007).
6.3
Mechanical Properties
Alginate gels and microbeads are largely used for food and biomedical applications and consequently they should have suitable mechanical properties to withstand the stresses (compression and shear) exerted. Therefore, it is highly desirable to clearly understand the molecular features controlling the stiffness and elasticity of the network to be able to select the proper alginate sample for different applications. From the theoretical point of view, the elasticity of a network is a measure of its deformation under stress and it is often described by the rubber elasticity theory. The classical rubber elasticity theory was derived for chemical networks in which the energy-at-break of the cross-links is so large that only the (entropy-
38
I. Donati and S. Paoletti 1000
900
diameter (μm)
800
700
600
500
400
300 0
1
2
3
4
5
Saline shifts
Fig. 16 Dimensional stability of alginate beads obtained from L. hyperborea alginate (FG = 0.66) (squares) and from an epimerized alginate (FG = 0.64) (circles). The lines are drawn to guide the eye. (Adapted from Mørch et al. 2007)
dominated) behavior of the polymer stretches connecting them needs to be considered to explain and describe elasticity. In addition, this theory describes the cross-links connecting the elastically active networks as stable and pointlike. Upon gel deformation, the free-energy increase of the network is caused by a decrease of entropy. According to the rubber elasticity approach, the shear modulus (G) can be related to the number of elastically active chains per unit volume (n), G = nkT =
cRT , Mc
(20)
or, if loose polymer ends that do not contribute to the network elasticity are taken into consideration, G=
c RT Mc
⎛ 2 Mc ⎞ ⎜⎝1 − ⎟, Mw ⎠
(21)
where k is Boltzmann’s constant, T the absolute temperature, c the weight concentration of the polymer, R the gas constant, Mc the average molecular weight of an — elastic chain, and Mw the weight-average molecular weight of the uncrosslinked polymer. The applicability of the rubber elasticity theory in the case of physical gels (including biopolymer gels) has been repeatedly questioned. In fact, even in the
Material Properties of Alginates
39
presence of cooperative sequences of interchain contacts held together by weak forces, the overall energy-at-break of the gel junctions may become comparable to the elastic (free) energy of the rubberlike component. Alginate gels certainly do not strictly meet the conditions required by the rubber elasticity theory. In fact, crosslinks are reversible junctions rather than permanent pointlike contacts and the chains between the junctions are restricted in their movements (although topological restrictions do not, per se, exclude the feasibility of a purely entropic description of the network; Gardel et al. 2004). Therefore, a purely entropic approach certainly does not capture all the physics of these systems: rather, both entropic and enthalpic terms contribute to the elasticity of alginate gels. The enthalpic contribution is negligible at small deformations (Doi and Kuzuu 1980; Higgs and Ball 1989) but it becomes important when large strains are exerted on the hydrogel. In contrast, in the small deformation range, where the linearity between s and g is fulfilled, the entropic contribution is expected to prevail (Higgs and Ball 1989; Linke et al. 1998). In general, the rheological characterization of alginate gels is performed to quantify the apparent modulus of rigidity based on small deformations in the longitudinal direction (compression or elongation) or on small-shear deformation. The first method provides Young’s modulus, E, while the latter gives the shear modulus, G, where G=
E , 2 (1 + r )
(22)
with r the Poisson ratio (0.5 for incompressible materials). The viscoelastic nature of the alginate-based networks, like the other polysaccharidebased systems, can be explored by dynamic measurements which allow determination of the elastic (G¢) and loss (G¢¢) moduli. Alginate hydrogels are cold-setting physical networks and their setting is independent of temperature (see Section 5). However, since alginate networks are considered to be nonequilibrium gels, the kinetics of gel setting, as well as the properties of the final system, will depend on the thermal history. The sol–gel transition in the case of alginate gels is accomplished by addition of divalent ions such as calcium. Several models have been proposed to describe the cross-linking of polymers at their critical gel point and the one formulated by Winter and Chambron (Winter and Chambon 1986; Chambon and Winter 1987) was found to be applicable in the case of alginate gels (Liu et al. 2003; Lu et al. 2005, 2006). In particular, a power-law dependence of the storage and loss moduli on the angular frequency was predicted to hold at the critical gel point: G ′ (w ) ≈ G ′′ (w ) ≈ w n ,
(23)
where n is the critical gel exponent. In the case of alginate, n was found to be basically independent of the molecular weight for low molecular weight alginate samples, having an almost constant value of 0.71, which is in good agreement with the percolation theory predictions (de Gennes 1979; Stauffer 1985; Martin et al. 1988).
40
I. Donati and S. Paoletti
In contrast, for high molecular weight alginate chains n decreases upon increasing the polymer concentration (Lu et al. 2005). At the same time, the fractal dimension at the critical gelation point for the high molecular weight alginate shows a concentration dependence, while it is basically independent of concentration for the low molecular weight alginate, pointing to a more densely filled network in the former case.
6.3.1
Small Deformation Range
The small deformation range refers to the part of the mechanical deformation curve for alginate in which a linear relationship holds between the strain (e) on the network and the stress exerted (s) (Hooke’s law). The elastic modulus is frequently used to measure the gel strength of alginate networks. In particular, Young’s modulus (E) is easily calculated from the slope of the linear e–s curve recorded in uniaxial compression: F Δl = E ⇒ s = Eε. A l
(24)
From the technical point of view, the mechanical properties of alginate depend on the deformation rate, the extent of deformation, and eventual rupture of the gel (Mitchell and Blanshard 1976). In general, the elastic modulus is affected by the amount of cross-linking ion used, the concentration and composition of the alginate, and, to certain extent, its molecular weight. The dependence of the mechanical properties on the concentration of the cross-linking ion can be easily understood, as a higher number of junctions (and hence a higher number of elastically active chains per unit volume) will be formed when a higher number of divalent ions are provided to the system. For this specific reason, calcium-saturated alginate hydrogels are better performing than the calcium-limited ones. By the same means (i.e., number of elastically active chains per unit volume), the polysaccharide concentration will have a strong impact on the elastic modulus of the alginate gel. In particular, it has been reported that Young’s modulus of alginate gels scales with the polysaccharide concentration according to the power law E μ c2 (Smidsrød and Haug 1972). This correlation is particularly useful when syneresis occurs and the variation of the concentration of the polysaccharide in the gel has to be accounted for (Martinsen et al. 1989). Alginate gels made by calcium diffusion (saturated gels) showed a molecularweight dependence of the elastic modulus up to a molecular weight of the polysaccharide of approximately 100,000. In contrast, internally set gels show more pronounced molecular-weight dependence, with the elastic modulus being affected by the dimension of the chain up to a molecular weight of 300,000 (Draget et al. 1993). This has been correlated with the limited amount of calcium generally used for the preparation of the latter hydrogels, which would induce a higher presence of loose ends (nonelastically active chains) in the network (Draget et al. 2005).
Material Properties of Alginates
41
The type of divalent cation used for the gelation also has a notable impact on the mechanical properties of the alginate hydrogels (Smidsrød 1974; Mørch et al. 2006). It has been reported that the minimum length of G-blocks required for junction formation decreases upon increasing the affinity of the ions towards the alginate chains (Stokke et al. 1991). As a consequence, cations that bind strongly to the uronic moieties present in the polysaccharide form a higher number of junctions (and hence elastically active chains) per unit volume. This effect, combined with a higher strength of the junction itself, leads to an increase of the mechanical properties (Young’s modulus) of the hydrogel (Stokke et al. 1991). From a general point of view, the rigidity of alginate gels decreases in the order Pb>Cu,Ba>Sr>Cd>Ca>Ni>Zn>Co>Mn. The composition of the polysaccharide plays an important role in the mechanical properties of the hydrogels. In particular, the gel strength was found to correlate with the length of the G-blocks (NG>1) rather than with the overall content of the guluronic acid residues (fraction of guluronic acid units, FG). A profound effect on Young’s modulus was detected for NG>1 ranging from approximately 4 to 15 (Fig. 17). Additional evidence for the importance of the composition for the mechanical properties of alginate gel was recently reported. In fact, it was noted that the enzymatic elongation of alternating sequences (MG-blocks) in the polysaccharide chain leads to an increase of Young’s modulus, E. This effect was first attributed to the higher intrinsic flexibility of the MG-blocks with respect to the other blocks composing alginate (Draget et al. 2000). However, the possible involvement of alternating sequences in calcium-induced junctions, whether heteropolymeric MG/GG or homopolymeric MG/MG junctions, provides an additional explanation for their effect on E (Donati et al. 2005b). In addition, it has been reported that the different block structures could have different binding ability with respect to the divalent ions.
100
E (kPa)
80 60 40 20 0
0
4
8 NG>1
12
16
Fig. 17 Young’s modulus of alginate gels as a function of the average G-block length (NG>1). (Adapted and modified from Skjåk-Bræk et al. 1986b)
42
6.3.2
I. Donati and S. Paoletti
Large Deformation Range
The study of large deformation properties for alginate gels is undoubtedly highly correlated with their application in food engineering, where typical processes such as cutting/slicing and eating have to be considered. However, a clear understanding of the network deformation behavior certainly also has a deep impact in the development of alginate-based materials to be used as bioreactors for regenerative medicine, since the forces exerted onto the gel implants likely exceed the linear stress–strain response. It is commonly known that the stress-strain curve of alginate gels displays a linear region that extends approximately up to a deformation of 10–12% of the specimen. However, for most saturated alginate gels the stress–strain linearity holds up to a maximum deformation of approximately 6–8%. The linear region is followed by a nonlinear trend described as “strain-hardening,” which is a rather general phenomenon for biopolymer gels and consists of an upturn of the stress-strain curve. In addition, when calcium-saturated alginate gels are considered, a region of plastic behavior, where notable deformations are caused by a small increase in the stress applied, can be found (Fig. 18). It has been shown that the presence and extension of the “plastic region” is related to the presence of long alternating sequences within the alginate chain (Mørch et al. 2008). Several interpretations for the mechanism of the strain hardening of hydrogels have been proposed. Flory argued that the upturn of the stress–strain curve is mainly attributed to strain-induced crystallization. In contrast, Treloar (1949) proposed a
σtrue (kPa)
30
20
10
0 0.0
0.1
0.2
γH
0.3
0.4
0.5
Fig. 18 Dependence of the true stress [strue = s(1−e)] on the Hencky strain [gH = −ln(1−e)] for saturated hydrogels from L. hyperborea. The dotted line represents the continuation of the linear stress-strain relationship. (Donati I and Paoletti S, unpublished data)
Material Properties of Alginates
43
non-Gaussian network elasticity approach, thus based on the finite extensibility of the polymer chains, that adequately explained the nonlinear behavior. Different investigations have focused on the large deformation properties of alginate gels (Mitchell and Blanshard 1976; Mitchell 1980; Mancini et al. 1999a; Zhang et al. 2005, 2007). In all cases, the presence of ion-induced junctions – which are by far larger that the pointlike cross-links of the natural rubber networks – is considered the main factor responsible of the upturn of the stress-strain curve and in general for the large deformation properties of alginate networks. Zhang et al. (2007) have recently proposed the possible direct involvement of junctions in the strain hardening of alginate gels. According to their model, the overall stress, s, can be described as being composed of two contributions: s = sC + sJ ,
(25)
where sC is the stress arising from the (entropic) deformation of elastically active chains and sJ is the stress produced by the (energetic) deformation of the junctions. In practice, at small strains the physical (ion-induced) junctions will not be deformed and a purely entropic description of the system holds. In contrast, when a large strain is reached, the stress will be transmitted to the network junctions deforming them and, eventually, causing their breakage (Zhang et al. 2007). As such, the rodlike junction zones contribute to a deformation-dependent stress which is the main cause of the strain-hardening of the hydrogel. This latter contribution has been described by the mathematical model of the finite deformation elasticity of gels formed from stiff, rodlike polymer chains (Doi and Kuzuu 1980), which predicts a nonlinear stress-strain curve dominated by an energetic (enthalpic) contribution. In the case of an alginate network, the fracture gel stress, i.e., the stress required to cause network breakage, can be safely correlated with the strength and number of junctions per unit volume. However, there is no direct correlation between the rupture strength and the elastic modulus (Mitchell and Blanshard 1976; Mitchell 1980). In fact, it has been reported that hydrogels from M-rich alginates, despite their lower elastic modulus, typically show a higher rupture strength than those obtained from G-rich alginates. Therefore, M-rich alginates are recognized to be more elastic, while G-rich alginates form more brittle gels. The length of the junction is also correlated with the rupture strength, as it is easily expected that upon application of the stress the shorter the junction is, the lower is the stress required to break it. After the breakage of the junction, the stress will be transferred to the neighboring junctions, eventually accelerating gel failure. It is interesting to notice that for alginate gels the fracture stress depends upon both the divalent cation used and the polysaccharide concentration, while the fracture strain is not apparently affected by these two parameters. Moreover, alginate hydrogels being viscoelastic materials, the fracture characteristics are affected by the deformation rate (Zhang et al. 2005). However, the fracture stress, in contrast to the elastic modulus of the alginate gels, does not become constant above a certain molecular weight (Mitchell 1979).
44
I. Donati and S. Paoletti
An empirical characterization of the behavior of alginate gels at large deformations has been proposed recently (Mancini et al. 1999b). In particular, the stress (s)–strain (e) curves have been shown to be accurately reproduced by a power equation (Peleg and Campanella 1989): s = je p,
(26)
where j is the rigidity constant that represents a measure of stiffness, while p is the degree of concavity. When p = 1, Eq. 26 reduces to Hooke’s law (and j coincides with the modulus of elasticity), while for p >1 or p <1, an upward or downward concavity, respectively, is detected. In the case of alginate, n was found to be approximately 2.17 independent of the alginate composition and concentration (Mancini et al. 1999b). Equation 26 was also found to hold at the gel fracture point, while the work for rupture was found to be very well represented by the following relation (Mancini et al. 1999b): σR
LR = ∫ σ de = k 0
e Rn+1 , n +1
(27)
where sR and eR are the stress and strain at failure, respectively. Mitchell and Blanshard (1976) discussed the viscoelastic behavior of the alginate gels by means of creep compliance measurements. The experimental curves were satisfactorily modeled using a combination of one Maxwell and two Voigt elements. From such measurements, the authors concluded that the cross-links in the hydrogel were not permanent but, upon the application of sufficiently high stresses, displayed the tendency to move and, eventually, to break. These reorientations did not seem to be present when small deformations were applied, but rather they could be regarded as a characteristic of the large deformation range. Recently, creep compliance measurements (Fig. 19) were used to point to a possible role of long MG-block sequences in alginate gels. The experimental curves were modeled accurately by means of one Maxwell and one Voigt model and the Newtonian viscosity, hN, was obtained for alginate gels prepared under different experimental conditions (Eq. 28).
)
J (t ) = J 0 + J1 (1 − e − t / τ +
t , ηN
(28)
where J(t) is the measured compliance, J0 and J1 are the compliance values of the Maxwell and Voigt springs, respectively, and t is the retardation time associated with the Voigt element. It is to be noted that upon treatment of natural alginate samples with the AlgE4 epimerase, which causes an elongation of the MG-blocks, an increase of hN was detected. This effect was traced back to the presence of extended secondary MG/ MG junctions which, by bridging alternating sequences in the gel network via an ion-mediated mechanism, enhance the frictional forces between polymer chains, reducing their slipping (Donati et al. 2005b).
Material Properties of Alginates
45
14
12 ηN–1
J (Pa–1) * 105
10 J0
8
6
4
2
0
0
1000
2000 time (s)
3000
4000
Fig. 19 Example of viscoelastic behavior (creep compliance curve) for L. hyperborea alginate. Inset: The combination of one Maxwell and one Voigt model used to fit the experimental data. (Adapted from Donati et al. 2005b)
6.4
Diffusion Properties
Diffusion properties of alginate hydrogels are determined by the pore size and the pore size distribution within the network. It has been reported that calcium-treated alginate forms networks characterized by a pore size between 50 and 1,500Å (Andresen et al. 1977). Self-diffusion of small molecules is very little affected by the alginate hydrogel. In fact, the rates of self-diffusion of glucose and ethanol have been reported to be as high as about 90% of their corresponding values in water (Tanaka et al. 1984b). The diffusion of larger molecules, such as proteins, is somehow restricted, although proteins with molecular weight as high as 300,000 are able to diffuse out of the hydrogel with a rate that depends on their molecular size (Tanaka et al. 1984b; Martinsen et al. 1989). The diffusion of molecules through alginate hydrogels depends upon both the cross-linking ion concentration and the polymer concentration and composition. The use of a higher concentration of divalent ions, such as calcium, will induce the formation of a more compact gel with a possible collapse of some junctions. This leads to a pore size increase and thus greater diffusion of the protein (Gåserød et al. 1998). At variance, the replacement of Ca2+ ions by Ba2+ ions does not seem to lead to modifications in the permeability of the alginate hydrogels (Mørch et al. 2006).
46
I. Donati and S. Paoletti
An increase of the polysaccharide concentration induces a reduction in the pore size, while the increase in the content of the G residues in the polymer chain has the opposite effect. The latter effect has been correlated with the condensation of junctions upon the addition of the divalent ion which is favored by the presence of long G-blocks (Martinsen et al. 1989; Gåserød et al. 1998). The coating of alginate with a polycation, such as poly(l-lysine), chitosan and polyethyleneimine (Tanaka et al. 1984a; Kulseng et al. 1997; Gåserød et al. 1999), enhances the polymer density on the surface of the bead, thus limiting the diffusion of the proteins. In particular, a proper polycation coating of alginate bead leads to a selective capsule permeability, thus allowing the free diffusion of proteins as small as insulin, but preventing diffusion of antibodies through the capsule membrane. These results represent the basis for proposing alginate gel beads for the encapsulation of Langerhans islets for the treatment of type I diabetes.
References Andresen I-L, Skipnes O, Smidsrød O, Østgaard K, Hemmer PC (1977) Some biological functions of matrix components in benthic algae in relation to their chemistry and the composition of seawater. ACS Symp Ser 48:361–381 Arnott S, Bian W, Chandrasekaran R, Manis B (2000) Lessons for today and tomorrow from yesterday - the structure of alginic acid. Fibre Diffr Rev 36:44–51 Atkins ED, Nieduszynski IA, Mackie W, Parker KD, Smolko EE (1973a) Structural components of alginic acid. I. The crystalline structure of poly-b-D-mannuronic acid. Results of x-ray diffraction and polarized infrared studies. Biopolymers 12:1865–1878 Atkins ED, Nieduszynski IA, Mackie W, Parker KD, Smolko EE (1973b) Structural components of alginic acid. II. The crystalline structure of poly-alpha-L-guluronic acid. Results of x-ray diffraction and polarized infrared studies. Biopolymers 12:1879–1887 Atsuki K, Tomoda Y (1926) Studies on seaweeds of Japan I. The chemical constituents of Laminaria. J Soc Chem Ind Jpn 29:509–517 Benegas JC, Di Blas A, Paoletti S, Cesàro A (1992) Some aspects of the enthalpy of dilution of biological polyelectrolytes. J Therm Anal Calorim 38:2613–2620 Berth (1992) Methodical aspects of characterization of alginate and pectate by light scattering and viscometry coupled with GPC. Carbohydr Polym 19:1 Blasi P, Giovagnoli S, Schoubben A, Ricci M, Rossi C, Luca G, Basta G, Calafiore R (2006) Preparation and in vitro and in vivo characterization of composite microcapsules for cell encapsulation. Int J Pharm 324:27–36 Braccini I, Grasso RP, Pérez S (1999) Conformational and configurational features of acidic polysaccharides and their interactions with calcium ions: a molecular modeling investigation. Carbohydr Res 317:119–130 Braccini I, Pérez S (2001) Molecular basis of Ca2+-induced gelation in alginates and pectins: the egg-box model revisited. Biomacromolecules 2:1089–1096 Campos M, Martínez-Salazar JM, Lloret L, Moreno S, Nuñez C, Espín G, Soberón-Chavéz G (1996) Characterization of the gene coding for GDP-mannose dehydrogenase (algD) from Azotobacter vinelandii. J Bacteriol 178:1793–1799 Cesàro A, Delben F, Flaibani A, Paoletti S (1988a) The interaction of lead (II) with glycuronans: U.V. absorption and circular dichroism spectra. Carbohydr Res 181:13–21 Cesàro A, Delben F, Paoletti S (1988b) Interaction of divalent cations with polyuronates. J Chem Soc Faraday Trans 1 84:2573–2584
Material Properties of Alginates
47
Chambon F, Winter HH (1987) Linear viscoelasticity at the gel point of a crosslinking PDMS with imbalanced stoichiometry. J Rheol 31:683–697 Cottrell IW, Kovacs P (1980) Alginates. In: Crawford HB, Williams J (eds) Handbook of watersoluble gums and resins. McGraw-Hill, Auckland, pp 21–43 de Gennes PG (1979) Scaling concepts in polymer physics. Cornell University Press, Ithaca Delben F, Cesáro A, Paoletti S, Crescenzi V (1982) Monomer composition and acetyl content as main determinants of the ionization behavior of alginates. Carbohydr Res 100:C46–C50 Dentini M, Rinaldi G, Risica D, Barbetta A, Skjåk-Bræk G (2005) Comparative studies on solution characteristics of mannuronan epimerized by C-5 epimerases. Carbohydr Polym 59:489–499 de Vos P, Faas MM, Strand B, Calafiore R (2006) Alginate-based microcapsules for immunoisolation of pancreatic islets. Biomaterials 27:5603–5617 Doi M, Kuzuu NY (1980) Nonlinear elasticity of rodlike macromolecules in condensed state. J Polym Sci Polym Phys 18:409–419 Donati I, Vetere A, Gamini A, Skjåk-Bræk G, Coslovi A, Campa C, Paoletti S (2003a) Galactosesubstituted alginate: preliminary characterization and study of gelling properties. Biomacromolecules 4:624–631 Donati I, Gamini A, Skjåk-Bræk G, Vetere A, Campa C, Coslovi A, Paoletti S (2003b) Determination of the diadic composition of alginate by means of circular dichroism: a fast and accurate improved method. Carbohydr Res 338:1139–1142 Donati I, Coslovi A, Gamini A, Skjåk-Bræk G, Vetere A, Campa C, Paoletti S (2004) Galactosesubstituted alginate 2: conformational aspects. Biomacromolecules 5:186–196 Donati I, Draget KI, Borgogna M, Paoletti S, Skjåk-Bræk G (2005a) Tailor-made alginate bearing galactose moieties on mannuronic residues: selective modification achieved by a chemoenzymatic strategy. Biomacromolecules 6:88–98 Donati I, Holtan S, Mørch YA, Borgogna M, Dentini M, Skjåk-Bræk G (2005b) New hypothesis on the role of alternating sequences in calcium-alginate gels. Biomacromolecules 6:1031–1040 Donati I, Benegas JC, Paoletti S (2006a) Polyelectrolyte study of the calcium-induced chain association of pectate. Biomacromolecules 7:3439–3447 Donati I, Cesàro A, Paoletti S (2006b) Specific interactions versus counterion condensation. 1. Nongelling ions/polyuronate systems. Biomacromolecules 7:281–287 Donati I, Haug IJ, Scarpa T, Borgogna M, Draget KI, Skjåk-Bræk G, Paoletti S (2007) Synergistic effects in semidilute mixed solutions of alginate and lactose-modified chitosan (chitlac). Biomacromolecules 8:957–962 Draget KI, Østagaard K, Smidsrød O (1991) Homogeneous alginate gels: a technical approach. Carbohydr Polym 14:159–178 Draget KI, Simensen MK, Onsøyen E, Smidsrød O (1993) Gel strength of Ca-limited alginate gels made in situ. Hydrobiologia 260–261:563–565 Draget KI, Skjåk Bræk G, Smidsrød O (1994) Alginic acid gels: the effect of alginate chemical composition and molecular weight. Carbohydr Polym 25:31–38 Draget KI, Strand B, Hartmann M, Valla S, Smidsrød O, Skjåk-Bræk G (2000) Ionic and acid gel formation of epimerised alginates; the effect of AlgE4. Int J Biol Macromol 27:117–122 Draget KI, Gåserød O, Aune I, Andersen PO, Storbakken B, Stokke BT, Smidsrød O (2001) Effects of molecular weight and elastic segment flexibility on syneresis in Ca-alginate gels. Food Hydrocolloids 15:485–490 Draget KI, Smidsrød O, Skjåk-Bræk G (2005) Alginates form algae. In: Steinbuchel A, Rhee SK (eds) Polysaccharides and polyamides in the food industry: properties, production, and patents. Wiley, Weinheim, pp 1–30 Dusseault J, Leblond FA, Robitaille R, Jourdan G, Tessier J, Ménard M, Henley N, Hallé JP (2005) Microencapsulation of living cells in semi-permeable membranes with covalently cross-linked layers. Biomaterials 26:1515–1522 Dusseault J, Langlois G, Meunier MC, Ménard M, Perreault C, Hallé JP (2008) The effect of covalent cross-links between the membrane components of microcapsules on the dissemination of encapsulated malignant cells. Biomaterials 29:917–924
48
I. Donati and S. Paoletti
Ertesvåg H, Doseth B, Larsen B, Skjåk-Bræk G, Valla S (1994) Cloning and expression of an Azotobacter vinelandii mannuronan C-5-epimerase gene. J Bacteriol 176:2846–2853 Ertesvåg H, Valla S, Skjåk-Bræk G (1996) Genetics and Biosynthesis of alginates. Carbohydrates in Europe 14:14–18 Ertesvåg H, Høidal HK, Skjåk-Bræk G, Valla S (1998) The Azotobacter vinelandii mannuronan C-5-epimerase AlgE1 consists of two separate catalytic domains. J Biol Chem 273:30927–30932 Fisher FG, Dörfel H (1955) Die Polyuronsäuren der Braunalgen (Kohlenhydrate der Algen). Z Physiol Chem 302:186–203 Fixman M (1964) Polyelectrolytes: a fuzzy sphere model. J Chem Phys 41:3772–3778 Flory PJ (1953) Principles of polymer chemistry. Oxford University Press, Ithaca Gacesa P (1992) Enzymic degradation of alginates. Int J Biochem 24:545–552 Gamini A, Paoletti S, Zanetti F (1992) Chain rigidity of polyuronated: static light scattering of aqueous solutions of hyaluronate and alginate. In: Harding SE, Sattelle DB, Bloomfield VA (eds) Laser light scattering in biochemistry. Royal Society of Chemistry, Cambridge, pp 294–311 Gardel ML, Shin JH, MacKintosh FC, Mahadevan L, Matsudaira P, Weitz DA (2004) Elastic behavior of cross-linked and bundled actin networks. Science 304:1301–1305 Gåserød O, Smidsrød O, Skjåk-Bræk G (1998) Microcapsules of alginate-chitosan–I: a quantitative study of the interaction between alginate and chitosan. Biomaterials 19:1815–1825 Gåserød O, Sannes A, Skjåk-Bræk G (1999) Microcapsules of alginate-chitosan–II. A study of capsule stability and permeability. Biomaterials 20:773–783 Gimmestad M, Sletta H, Ertesvåg H, Bakkevig K, Jain S, Suh Sj, Skjåk-Bræk G, Ellingsen TE, Ohman DE, Valla S (2003) The Pseudomonas fluorescens AlgG protein, but not its mannuronan C-5-epimerase activity, is needed for alginate polymer formation. J Bacteriol 185:3515–3523 Gorin PAJ, Spencer JFT (1966) Exocellular alginic acid from Azotobacter Vinelandii. Can J Chem 44:993–998 Govan JR, Fyfe JA, Jarman TR (1981) Isolation of alginate-producing mutants of Pseudomonas fluorescens, Pseudomonas putida and Pseudomonas mendocina. J Gen Microbiol 125:217–220 Grant GT, Morris ER, Rees DA, Smith PJC, Thom D (1973) Biological interactions between polysaccharides and divalent cations: the egg-box model. FEBS Lett 32:195–198 Grasdalen H (1983) High-field, 1H-n.m.r. spectroscopy of alginate: sequential structure and linkage conformations. Carbohydr Res 118:255–260 Grasdalen H, Larsen B, Smidsrød O (1977) 13C-N.m.r. studies of alginate. Carbohydr Res 56:C11–C15 Grasdalen H, Larsen B, Smidsrød O (1979) A p.m.r. study of the composition and sequence of uronate residues in alginates. Carbohydr Res 68:23–31 Grasdalen H, Larsen B, Smisrød O (1981) 13C-n.m.r. studies of monomeric composition and sequence in alginate. Carbohydr Res 89:179–191 Hartmann M, Holm OB, Johansen GAB, Skjåk-Bræk G, Stokke BT (2002a) Mode of action of recombinant Azotobacter vinelandii mannuronan C-5 epimerases AlgE2 and AlgE4. Biopolymers 63:77–88 Hartmann M, Duun AS, Markussen S, Grasdalen H, Valla S, Skjåk-Bræk G (2002b) Timeresolved 1H and 13C NMR spectroscopy for detailed analyses of the Azotobacter vinelandii mannuronan C-5 epimerase reaction. Biochim Biophys Acta Gen Subj 1570:104–112 Hartmann M, Dentini M, Draget KI, Skjåk-Bræk G (2006) Enzymatic modification of alginates with the mannuronan C-5epimerase AlgE4 enhances their solubility at low pH. Carbohydr Polym 63:257–262 Hasse C, Bohrer T, Barth P, Stinner B, Cohen R, Cramer H, Zimmermann U, Rothmund M (2000) Parathyroid xenotransplantation without immunosuppression in experimental hypoparathyroidism: long-term in vivo function following microencapsulation with a clinically suitable alginate. World J Surg 24:1361–1366 Haug A (1959a) Ion exchange properties of alginate fractions. Acta Chem Scand 13:1250–1251 Haug A (1959b) Fractionation of alginic acid. Acta Chem Scand 13:601–603
Material Properties of Alginates
49
Haug, A (1964) Composition and properties of alginates. Dissertation, Norwegian Institute of Technology, Trondheim Haug A, Larsen B (1963) The solubility of alginate at low pH. Acta Chem Scand 17:1653–1662 Haug A, Smidsrød O (1965) Fractionation of alginate by precipitation with calcium and magnesium ions. Acta Chem Scand 19:1221–1226 Haug A, Smidsrød O (1967) Strontium-calcium selectivity of alginates. Nature 215:757 Haug A, Smidsrød O (1970) Selectivity of some anionic polymers for divalent metal ions. Acta Chem Scand 24:843–854 Haug A, Larsen B, Smidsrød O (1963) The degradation of alginates at different pH values. Acta Chem Scand 17:1466–1468 Haug A, Larsen B, Smidsrød O (1966) A study on the constitution of alginic acid by partial hydrolysis. Acta Chem Scand 20:183–190 Haug A, Myklestad S, Larsen B, Smidsrød O (1967a) Correlation between chemical structure and physical properties of alginates. Acta Chem Scand 21:768–778 Haug A, Larsen B, Smidsrød O (1967b) Alkaline degradation of alginate. Acta Chem Scand 21:2859–2870 Haug A, Larsen B, Smidsrød O (1967c) Studies on the sequence of uronic acid residues in alginic acid. Acta Chem Scand 21:691–704 Heinemann M, Meinberg H, Büchs J, Koß HJ, Ansorge-Schumacher MB (2005) Method for quantitative determination of spatial polymer distribution in alginate beads using Raman spectroscopy. Appl Spectrosc 59:280–285 Hertzberg S, Moen E, Vogelsang C, Østgaard K (1995) Mixed photo-cross-linked polyvinyl alcohol and calcium-alginate gels for cell entrapment. Appl Microbiol Biot 43:10–17 Higgs PG, Ball RC (1989) Some ideas concerning the elasticity of biopolymer networks. Macromolecules 22:2432–2437 Hoffman AS (2002) Hydrogels for biomedical applications. Adv Drug Deliv Rev 54:3–12 Høidal HK, Ertesvåg H, Skjåk-Bræk G, Stokke BT, Valla S (1999) The recombinant Azotobacter vinelandii mannuronan C-5-epimerase AlgE4 epimerizes alginate by a nonrandom attack mechanism. J Biol Chem 274:12316–12322 Holtan S, Bruheim P, Skjåk-Bræk G (2006) Mode of action and subsite studies of the guluronan block-forming mannuronan C-5 epimerases AlgE1 and AlgE6. Biochem J 395:319–329 Indergaard M, Skjåk-Bræk G (1987) Characteristics of alginate from Laminaria digitata cultivated in a high-phosphate environment. Hydrobiologia 151–152:541–549 Joki T, Machluf M, Atala A, Zhu J, Seyfried NT, Dunn IF, Abe T, Carroll RS, Black PM (2001) Continuous release of endostatin from microencapsulated engineered cells for tumor therapy. Nat Biotechnol 19:35–39 Keith RC, Keith LMW, Hernandez-Guzman G, Uppalapati SR, Bender CL (2003) Alginate gene expression by Pseudomonas syringae pv. tomato DC3000 in host and non-host plants. Microbiology 149:1127–1138 Kohn R, Luknar O (1977) Intermolecular calcium ion binding on polyuronates - polygalacturonate and polyguluronate. Collect Czech Chem C 42:731–744 Kong HJ, Smih MK, Mooney DJ (2003) Designing alginate hydrogels to maintain viability of immobilized cells. Biomaterials 24:4023–4029 Kulseng B, Thu B, Espevik T, Skjåk-Bræk G (1997) Alginate polylysine microcapsules as immune barrier: permeability of cytokines and immunoglobulins over the capsule membrane. Cell Transplant 6:387–394 Larsen B, Painter TJ (1969) The periodate-oxidation limit of alginate. Carbohydr Res 10:186–187 Larsen B, Smidsrød O, Painter T, Haug A (1970) Calculation of the nearest-neighbour frequencies in fragments of alginate from the yields of free monomer after partial hydrolysis. Acta Chem Scand 24:726–728 Lawrie G, Keen I, Drew B, Chandler-Temple A, Rintoul L, Fredericks P, Grøndahl L (2007) Interactions between alginate and chitosan biopolymers characterized using FTIR and XPS. Biomacromolecules 8:2533–2541 Lee KY, Mooney DJ (2001) Hydrogels for tissue engineering. Chem Rev 101:1869–1880
50
I. Donati and S. Paoletti
Lim F, Sun AM (1980) Microencapsulated islets as bioartificial endocrine pancreas. Science 210:908–910 Linke WA, Ivemeyer M, Mundel P, Stockmeier MR, Kolmerer B (1998) Nature of PEVK-titin elasticity in skeletal muscle. Proc Natl Acad Sci U S A 95:8052–8057 Linker A, Evans LR (1984) Isolation and characterization of an alginase from mucoid strains of Pseudomonas aeruginosa. J Bacteriol 159:958–964 Linker A, Jones RS (1966) A new polysaccharide resembling alginic acid isolated from pseudomonads. J Biol Chem 241:3845–3851 Littlecott GW (1982) Food gels – the role of alginates. Food Technol Aust 34:412–418 Liu X, Qian L, Shu T, Tong Z (2003) Rheology characterization of sol-gel transition in aqueous alginate solutions induced by calcium cations through in situ release. Polymer 44:407–412 Lu L, Liu X, Dai L, Tong Z (2005) Difference in concentration dependence of relaxation critical exponent n for alginate solutions at sol-gel transition induced by calcium cations. Biomacromolecules 6:2150–2156 Lu L, Liu X, Tong Z, Gao Q (2006) Critical exponents and self-similarity for sol-gel transition in aqueous alginate systems induced by in situ release of calcium cations. J Phys Chem B 110: 25013–25020 Mackie W, Noy R, Sellen DB (1980) Solution properties of sodium alginate. Biopolymers 19:1839–1860 Mancini M, Moresi M, Rancini R (1999a) Uniaxial compression and stress relaxation tests on alginate gels. J Texture Stud 30:639–657 Mancini M, Moresi M, Rancini R (1999b) Mechanical properties of alginate gels: empirical characterisation. J Food Eng 39:369–378 Manning GS (1969a) Limiting laws and counterion condensation in polyelectrolyte solutions I. Colligative properties. J Chem Phys 51:924–933 Manning GS (1969b) Limiting laws and counterion condensation in polyelectrolyte solutions II. Self-Diffusion of the small ions. J Chem Phys 51:934–938 Manning GS (1977) Limiting laws and counterion condensation in polyelectrolyte solutions: IV. The approach to the limit and the extraordinary stability of the charge fraction. Biophys Chem 7:95–102 Manning GS (1978) Limiting laws and counterion condensation in polyelectrolyte solutions: V. Further development of the chemical model. Biophys Chem 9:65–70 Martin JE, Adolf D, Wilcoxon JP (1988) Viscoelasticity of near-critical gels. Phys Rev Lett 61:2620 Martinsen A, Skjåk Bræk G, Smidsrød O (1989) Alginate as immobilization material: I. Correlation between chemical and physical properties of alginate gel beads. Biotechnol Bioeng 33:79–89 Martinsen A, Skjåk-Bræk G, Smidsrød O, Zanetti F, Paoletti S (1991) Comparison of different methods for determination of molecular weight and molecular weight distribution of alginates. Carbohydr Polym 15:171–193 Martinsen A, Storrø I, Skjårk-Bræk G (1992) Alginate as immobilization material: III. Diffusional properties. Biotechnol Bioeng 39:186–194 Mitchell JR (1979) Rheology of polysaccharide solutions and gels. In: Blanshard JMV, Mitchell JR (eds) Polysaccharides in food. Butterworths, London, pp 51–72 Mitchell JR (1980) The rheology of gels. J Texture Stud 11:315–337 Mitchell JR, Blanshard JMV (1976) Rheological properties of alginate gels. J Texture Stud 7:219–234 Moe ST, Skjåk-Bræk G, Elgsaeter A, Smidsrød O (1993) Swelling of covalently crosslinked alginate gels: influence of ionic solutes and nonpolar solvents. Macromolecules 26:3589–3597 Mørch YA, Donati I, Strand BL, Skjåk-Bræk G (2006) Effect of Ca2+, Ba2+, and Sr2+ on alginate microbeads. Biomacromolecules 7:1471–1480 Mørch YA, Donati I, Strand BL, Skjåk-Bræk G (2007) Molecular engineering as an approach to design new functional properties of alginate. Biomacromolecules 8:2809–2814 Mørch YA, Holtan S, Donati I, Strand BL, Strand WI, Skjåk-Bræk G (2009) Mechanical properties of C5-epimerised alginates. Biomacromolecules 9:2360–2368
Material Properties of Alginates
51
Morris ER, Rees DA, Sanderson GR, Thom D (1975) Conformation and circular dichroism of uronic acid residues in glycosides and polysaccharides. J Chem Soc Perkin Trans 2 1418–1425 Morris ER, Rees DA, Thom D, Boyd J (1978) Chiroptical and stoichiometric evidence of a specific, primary dimerisation process in alginate gelation. Carbohydr Res 66:145–154 Morris ER, Rees DA, Thom D (1980) Characterisation of alginate composition and blockstructure by circular dichroism. Carbohydr Res 81:305–314 Morris ER, Cutler AN, Ross-Murphy SB, Rees DA, Price J (1981) Concentration and shear rate dependence of viscosity in random coil polysaccharide solutions. Carbohydr Polym 1:5–21 Odijk T, Houwaart AC (1978) On the theory of the excluded-volume effect of a polyelectrolyte in a 1–1 electrolyte solution. J Polym Sci Polym Phys 16:627–639 Painter TJ (1983) Algal polysaccharides. In: Aspinall GO (ed) The polysaccharides. Academic, New York, pp 196–286 Painter T, Smidsrød O, Larsen B, Haug A (1968) A computer study of the changes in compositiondistribution occurring during random depolymerization of a binary linear heteropolysaccharide. Acta Chem Scand 22:1637–1648 Paoletti S, Cesàro A, Delben F, Crescenzi V, Rizzo R (1985) Polyelectrolytic aspects of conformational transitions and interchain interactions in ionic polysaccharide solutions: comparison of theory and microcalorimetric data. In: Dubin PL (ed) Microdomains in polymer solutions. Plenum, New York, pp 159–189 Pedersen SS, Kharazmi A, Espersen F, Hoiby N (1990) Pseudomonas aeruginosa alginate in cystic fibrosis sputum and the inflammatory response. Infect Immun 58:3363–3368 Peleg M, Campanella OH (1989) The mechanical sensitivity of soft compressible testing machines. J Rheol 33:455–467 Rokstad AM, Holtan S, Strand B, Steinkjer B, Ryan L, Kulseng B, Skjåk-Bræk G, Espevik T (2002) Microencapsulation of cells producing therapeutic proteins: optimizing cell growth and secretion. Cell Transplant 11:313–324 Rokstad AM, Donati I, Borgogna M, Oberholzer J, Strand BL, Espevik T, Skjåk-Bræk G (2006) Cell-compatible covalently reinforced beads obtained from a chemoenzymatically engineered alginate. Biomaterials 27:4726–4737 Rowley JA, Mooney DJ (2002) Alginate type and RGD density control myoblast phenotype. J Biomed Mater Res 60:217–223 Rowley JA, Madlambayan G, Mooney DJ (1999) Alginate hydrogels as synthetic extracellular matrix materials. Biomaterials 20:45–53 Schimdt E, Vocke F (1926) Zur Kenntnis der Polyglykuronsäuren. Chem Ber 59:1585–1588 Sherbrock-Cox V, Russell NJ, Gacesa P (1984) The purification and chemical characterisation of the alginate present in extracellular material produced by mucoid strains of Pseudomonas aeruginosa. Carbohydr Res 135:147–154 Sikorski P, Mo F, Skjåk-Bræk G, Stokke BT (2007) Evidence for egg-box-compatible interactions in calcium-alginate gels from fiber X-ray diffraction. Biomacromolecules 8:2098–2103 Sime W (1990) Alginates. In: Harris P (ed) Food gels. Elsevier, London, pp 53–78 Skjåk-Bræk G, Larsen B, Grasdalen H (1985) The role of O-acetyl groups in the biosynthesis of alginate by Azotobacter vinelandii. Carbohydr Res 145:169–174 Skjåk-Bræk G, Grasdalen H, Larsen B (1986a) Monomer sequence and acetylation pattern in some bacterial alginates. Carbohydr Res 154:239–250 Skjåk-Bræk G, Smidsrød O, Larsen B (1986b) Tailoring of alginates by enzymatic modification in vitro. Int J Biol Macromol 8:330–336 Skjåk-Bræk G, Grasdalen H, Smidsrød O (1989a) Inhomogeneous polysaccharide ionic gels. Carbohydr Polym 10:31–54 Skjåk-Bræk G, Zanetti F, Paoletti S (1989b) Effect of acetylation on some solution and gelling properties of alginates. Carbohydr Res 185:131–138 Smetana K (1993) Cell biology of hydrogels. Biomaterials 14:1046–1050 Smidsrød O (1970) Solution properties of alginate. Carbohydr Res 13:359–372 Smidsrød O (1974) Molecular basis of some physical properties of alginates in the gel state. Faraday Discuss Chem Soc 57:263–274
52
I. Donati and S. Paoletti
Smidsrød O, Draget KI (1996) Alginate gelation technology. In: Food colloids – proteins, lipids and polysaccharides. The Royal Society of Chemistry, pp 279–293 Smidsrød O, Haug A (1971) Estimation of the relative stiffness of the molecular chain in polyelectrolytes from measurements of viscosity at different ionic strengths. Biopolymers 10:1213–1227 Smidsrød O, Haug A (1972) Properties of poly(1,4-hexuronates) in the gel state. II. Comparison of gels of different chemical composition. Acta Chem Scand 26:79–88 Smidsrød O, Painter T (1973) Effect of periodate oxidation upon the stiffness of the alginate molecule in solution. Carbohydr Res 26:125–132 Smidsrød O, Skjåk-Bræk G (1990) Alginate as immobilization matrix for cells. Trends Biotechnol 8:71–78 Smidsrød O, Whittington SG (1969) Monte Carlo investigation of chemical inhomogeneity in copolymers. Macromolecules 2:42–44 Smidsrød O, Haug A, Larsen B (1963a) Degradation of alginate in the presence of reducing compounds. Acta Chem Scand 17:2628–2637 Smidsrød O, Haug A, Larsen B (1963b) The influence of reducing substances on the rate of regradation of alginates. Acta Chem Scand 17:1473–1474 Smidsrød O, Haug A, Larsen B (1966) The influence of pH on the rate of hydrolysis of acidic polysaccharides. Acta Chem Scand 20:1026–1034 Smidsrød O, Haug A, Larsen B (1967) Oxidative-reductive depolymerization: a note on the comparison of degradation rates of different polymers by viscosity measurements. Carbohydr Res 5:482–485 Smidsrød O, Haug A, Lian B (1972) Properties of poly(1,4-hexuronates) in the gel state. I. Evaluation of a method for the determination of stiffness. Acta Chem Scand 26:71–78 Smidsrød O, Glover RM, Whittington SG (1973) The relative extension of alginates having different chemical composition. Carbohydr Res 27:107–118 Soon-Shiong P, Heintz RE, Merideth N, Yao QX, Yao Z, Zheng T, Murphy M, Moloney MK, Schmehl M, Harris M, Mendez R, Mendez R, Sandford PA (1994) Insulin independence in a type 1 diabetic patient after encapsulated islet transplantation. Lancet 343:950–951 Stanford ECC (1881) Br Patent 142 Stanford ECC (1883a) New substance obtained from some of the commoner species of marine algae; algin. Chem News 47:254–257 Stanford ECC (1883b) New substance obtained from some of the commoner species of marine algae; algin. Chem News 47:267–269 Stauffer D (1985) Introduction to percolation theory. Taylor and Francis, London Steginsky CA, Beale JM, Floss HG, Mayer RM (1992) Structural determination of alginic acid and the effects of calcium binding as determined by high-field n.m.r. Carbohydr Res 225:11–26 Stokke BT, Smidsrød O, Bruheim P, Skjåk-Bræk G (1991) Distribution of uronate residues in alginate chains in relation to alginate gelling properties. Macromolecules 24:4637–4645 Stokke BT, Smidsrød O, Brant DA (1993) Predicted influence of monomer sequence distribution and acetylation on the extension of naturally occurring alginates. Carbohydr Polym 22:57–66 Stokke BT, Draget KI, Smidsrød O, Yuguchi Y, Urakawa H, Kajiwara K (2000) Small-angle X-ray scattering and rheological characterization of alginate gels. 1. Ca-alginate gels. Macromolecules 33:1853–1863 Strand BL, Gåserød O, Kulseng B, Espevik T, Skjåk-Bræk G (2002) Alginate-polylysine-alginate microcapsules: effect of size reduction on capsule properties. J Microencapsul 19:615–630 Strand BL, Mørch YA, Espevik T, Skjåk-Bræk G (2003) Visualization of alginate-poly-L-lysinealginate microcapsules by confocal laser scanning microscopy. Biotechnol Bioeng 82:386–394 Svanem BIG, Skjåk-Bræk G, Ertesvåg H, Valla S (1999) Cloning and expression of three new azotobacter vinelandii genes closely related to a previously described gene family encoding mannuronan C-5-epimerases. J Bacteriol 181:68–77 Tanaka H, Kurosawa H, Kokufuta E (1984a) Preparation of immobilized glucoamylase using Ca-alginate gel coated with partially quaternized poly(ethyleneimine). Biotechnol Bioeng 26: 1393–1394
Material Properties of Alginates
53
Tanaka H, Matsumura M, Veliky IA (1984b) Diffusion characteristics of substrates in Ca-alginate gel beads. Biotechnol Bioeng 26:53–58 Taqieddin E, Amiji M (2004) Enzyme immobilization in novel alginate-chitosan core-shell microcapsules. Biomaterials 25:1937–1945 Thu B, Bruheim P, Espevik T, Smidsrød O, Soon-Shiong P, Skjåk-Bræk G (1996a) Alginate polycation microcapsules: I. Interaction between alginate and polycation. Biomaterials 17:1031–1040 Thu B, Bruheim P, Espevik T, Smidsrød O, Soon-Shiong P, Skjåk-Bræk G (1996b) Alginate polycation microcapsules: II. Some functional properties. Biomaterials 17:1069–1079 Thu B, Gåserød O, Paus D, Mikkelsen A, Skjåk-Bræk G, Toffanin R, Vittur F, Rizzo R (2000) Inhomogeneous alginate gel spheres: an assessment of the polymer gradients by synchrotron radiation-induced X-ray emission, magnetic resonance microimaging, and mathematical modeling. Biopolymers 53:60–71 Tielen P, Strathmann M, Jaeger KE, Flemming HC, Wingender J (2005) Alginate acetylation influences initial surface colonization by mucoid Pseudomonas aeruginosa. Microbiol Res 160:165–176 Toft K, Grasdalen H, Smidsrød O (1986) Synergistic gelation of alginates and pectins. ACS Symp Ser 310:117–132 Treloar LRJ (1949) The physics of rubber elasticity, 3rd edn. Clarendon, Oxford Vold IMN, Kristiansen KA, Christensen BE (2006) A study of the chain stiffness and extension of alginates, in vitro epimerized alginates, and periodate-oxidized alginates using size-exclusion chromatography combined with light scattering and viscosity detectors. Biomacromolecules 7:2136–2146 Vold IMN, Kristiansen KA, Christensen BE (2007) A study of the chain stiffness and extension of alginates, in vitro epimerized alginates, and periodate-oxidized alginates using sizeexclusion chromatography combined with light scattering and viscosity detectors (addition/ correction). Biomacromolecules 8:2627 Winn SR, Tresco PA, Zielinski B, Greene LA, Jaeger CB, Aebischer P (1991) Behavioral recovery following intrastriatal implantation of microencapsulated PC12 cells. Exp Neurol 113:322–329 Winter HH, Chambon F (1986) Analysis of linear viscoelasticity of a crosslinking polymer at the gel point. J Rheol 30:367–382 Wong TY, Preston LA, Schiller NL (2000) Alginate lyase: review of major sources and enzyme characteristics, structure-function analysis, biological roles, and applications. Annu Rev Microbiol 54:289–340 Yang J, Goto M, Ise H, Cho CS, Akaike T (2002) Galactosylated alginate as a scaffold for hepatocytes entrapment. Biomaterials 23:471–479 Yilmazer G, Kokini JL (1991) Effect of polysorbate-60 on the stability of O/W emulsions stabilized by propylene glycol alginate and xanthan gum. J Texture Stud 22:289–301 Yilmazer G, Carrillo AR, Kokini JL (1991) Effect of propylene glycol alginate and xanthan gum on stability of O/W emulsions. J Food Sci 56:513–517 Yu J, Peñaloza-Vázquez A, Chakrabarty AM, Bender CL (1999) Involvement of the exopolysaccharide alginate in the virulence and epiphytic fitness of Pseudomonas syringae pv. syringae. Mol Microbiol 33:712–720 Zhang J, Daubert CR, Foegeding EA (2005) Fracture analysis of alginate gels. J Food Sci 70:e425–e431 Zhang J, Daubert CR, Foegeding EA (2007) A proposed strain-hardening mechanism for alginate gels. J Food Eng 80:157–165
Alginate Production: Precursor Biosynthesis, Polymerization and Secretion Bernd H.A. Rehm
Contents 1 2
Introduction.......................................................................................................................... Alginate Precursor Biosynthesis .......................................................................................... 2.1 Alginate Biosynthesis Genes ...................................................................................... 2.2 GDP-Mannuronic Acid Biosynthesis ......................................................................... 3 Alginate Polymerization and Secretion ............................................................................... 3.1 Alginate Polymerization ............................................................................................. 3.2 Alginate Export ........................................................................................................... 4 Conclusion and Future Perspectives .................................................................................... References ..................................................................................................................................
56 57 58 59 61 61 65 67 67
Abstract The activated intracellular precursor of alginate biosynthesis is GDPmannuronic acid. Carbon sources are oxidized to acetyl-CoA, which enters the citric acid cycle, providing the intermediate oxaloacetate. Oxaloacetate is converted via gluconeogenesis into fructose 6-phosphate. The central metabolite fructose 6-phosphate derived from gluconeogenesis is then converted to the activated alginate precursor GDP-mannuronic acid employing alginate-specific biosynthesis enzymes. This conversion requires four enzymatic steps catalysed by bifunctional phosphomannose isomerase:GDP-mannose pyrophosphorylase, phosphomannomutase and GDP-mannose dehydrogenase. GDP-mannuronic acid is polymerized to alginate by a membrane-anchored glycosyltransferase which is presumably represented by Alg8, which has been suggested to be a subunit of a multiprotein complex spanning the cytoplasmic membrane (Alg44), the periplasm (AlgX, AlgK, AlgG, AlgL) and the outer membrane (AlgE). These periplasmic proteins have been proposed to form B.H.A. Rehm Institute of Molecular BioSciences, Massey University, Private Bag 11222, Palmerston North, New Zealand e-mail: [email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_2, © Springer-Verlag Berlin Heidelberg 2009
55
56
B.H.A. Rehm
a scaffold guiding the nascent alginate chain through the periplasm to the alginatespecific channel protein AlgE in the outer membrane. AlgE has been suggested to facilitate export of alginate through the outer membrane. The extended periplasmic C terminus of membrane-anchored Alg44 shows similarities to membrane fusion proteins and might colocalize the alginate polymerase (Alg8) with the export protein AlgE. The cytoplasmic N-terminal loop of Alg44 comprises a PilZ domain required for binding of the secondary messenger bis-(3¢–5¢)-cyclic dimeric guanosine monophosphate, the binding of which was found to be required for alginate production
1
Introduction
Alginates represent a family of non-repeating unbranched exopolysaccharides composed of various amounts of (1–4)-linked b-d-mannuronic acid and its C5-epimer a-l-guluronic acid. These sugar acids are distributed in blocks of continuous b-d-mannuronic acid residues (M-blocks), a-l-guluronic acid residues (G-blocks), or as alternating residues (MG-blocks) (Rehm and Valla 1997; Rehm 2005b). Alginates are synthesized by brown seaweeds and by bacteria belonging to the genera Pseudomonas and Azotobacter (Rehm 2002). The distribution of comonomer residues and in particular the presence of G-blocks were found to be similar in algal alginates and in alginates derived from Azotobacter vinelandii, whereas alginates from pseudomonads are different in that they lack G-blocks (Skjak-Braek et al. 1986). These structural differences lead to different material properties reflecting the different biological functions of the alginates (see “Material properties of alginates” by Donati and Paoletti, this volume). In brown algae and Azotobacter cysts (a dormant stage) alginate serves as a cell wall constituent, whereas in Pseudomonas it contributes to the biofilm matrix. In the opportunistic human pathogen Pseudomonas aeruginosa alginate is considered a virulence factor contributing to the formation of persistent biofilms after infection of the lungs of cystic fibrosis (CF) patients (Kobayashi 2005). Knowledge of alginate biosynthesis is most advanced in P. aeruginosa and hence in this chapter I will focus on this bacterium. The cytosolic biosynthesis steps leading to the formation of the activated precursor GDP-mannuronic acid have been extensively investigated and are well understood; however, the actual polymerization and secretion of alginate is poorly understood. Recent studies suggested the formation of a multiprotein complex spanning the cytoplasmic membrane, the periplasm as well as the outer membrane (Jain and Ohman 2005; Oglesby et al. 2008; Remminghorst and Rehm 2006b, c). Future research dedicated to unravelling these final steps of alginate production will be motivated by medical interest with respect to the design of alginate biosynthesis inhibitors as well as by biotechnological applications implementing the engineering of improved bacterial production organisms. Bacterial fermentation using production organisms obtained by metabolic engineering approaches should enable the biotechnological production of a wide range of defined alginates suitable for high-value applications as biomaterials in medicine (Rehm 2005; Rehm 2009). The unique material properties of alginates have already led to a variety of industrial applications, such
Alginate Production: Precursor Biosynthesis
57
as stabilizing, thickening and gelling agents in food production and immobilization of cells in pharmaceutical and biotechnology industries (Paul et al. 1986). Commercial alginates are currently exclusively produced from brown seaweeds.
2
Alginate Precursor Biosynthesis
The alginate precursor GDP-mannuronic acid is synthesized in the cytosol by stepwise conversion of the central metabolite fructose 6-phosphate (Fig. 1). Three alginate-specific enzymes catalyse the four biosynthesis steps and these enzymes
fructose-6-phosphate AlgA mannose-6-phosphate
acetyl-CoA
eogen
PDH
glucon
KDPG
esis
Carbon source
AlgC mannose-1-phosphate AlgA GDP-mannose AlgD
GDP-mannuronic acid TCA Cycle
Polymerization Modification Export
Alginate Fig. 1 Biosynthesis route of alginate in bacteria. KDPG ketodeoxyphosphogluconate pathway (Entner–Doudoroff pathway), PDH pyruvate dehydrogenase, AlgA phosphomannose isomerase– GDP-mannose pyrophosphorylase, AlgC phosphor-mannomutase, AlgD GDP-mannose dehydrogenase. Boxed intermediates are precursors of alginate. Dashed arrows indicate unknown biosynthesis steps
58
B.H.A. Rehm Multiprotein complex with polymerisation and export function
P P
DD
88
44 44
KK
EE
G G
XX
LL
II
JJ
FF
AA
Alginate modifying enzymes Precursor biosynthesis
P
C C
Fig. 2 The alginate biosynthesis operon and functional assignment of the genes
have been extensively characterized. Two of the genes (algD, algA) encoding these enzymes are localized in the alginate biosynthesis gene cluster and one gene (algC) is not colocalized (Fig. 2).
2.1
Alginate Biosynthesis Genes
The first alginate biosynthesis genes were discovered in P. aeruginosa motivated by the medical significance of this bacterium as an opportunistic human pathogen which often establishes chronic infections in the lung of CF patients (May et al. 1991). Alginate as a virulence factor contributes to the establishment of persistent biofilms after infection (Gacesa 1998; Russell and Gacesa 1988). The alginate biosynthesis genes and their function have been compiled in a recent review (Rehm 2002). Seminal research by Darzins and Chakrabarty (1984) created the foundation for the identification and analysis of alginate biosynthesis genes in P. aeruginosa. The whole genome sequence of P. aeruginosa is now available and has already enabled functional genomic studies with respect to alginate biosynthesis (Lizewski et al. 2004; Stover et al. 2000). At least 24 genes were found to be directly involved in alginate biosynthesis in P. aeruginosa, with a few genes encoding proteins not exclusively involved in alginate biosynthesis (Table 1) (Gaona et al. 2004; Goldberg et al. 1993; Ledgham et al. 2003; Lizewski et al. 2004; Olvera et al. 1999; Wu et al. 2004; Ye et al. 1994). Twelve biosynthesis genes are colocalized in the alginate biosynthesis gene cluster, the transcription of which is regulated by one promoter upstream of the algD gene (Fig. 2) (Chitnis and Ohman 1993). How alginate biosynthesis gene transcription is regulated is discussed in “Alginate gene regulation” by Ohman in this volume.
Alginate Production: Precursor Biosynthesis
59
Table 1 The alginate biosynthesis genes from Pseudomonas aeruginosa Gene
Gene product/function or putative functiona
algD Alg8 Alg44 algK algE algG algX
GDP-mannose dehydrogenase Glycosyltransferase/polymerase-export function? c-di-GMP binding-activation/membrane fusion protein? Periplasmic protein/multiprotein complex assembly Outer-membrane porin/alginate export? Mannuronan C-5-epimerase/biosynthesis? Periplasmic protein with high sequence similarity to algJ/scaffold protein sequestering MucD algL Alginate lyase/biosynthesis? PA1167 Alginate lyase (polyguluronate lyase)/biosynthesis? algI O-Acetylation algJ O-Acetylation algF O-Acetylation algA Phosphomannose isomerase/GDP-mannose pyrophosphorylase algB Member of ntrC subclass of two-component transcriptional regulators (cognate sensor kinase is KinB) algC Phosphomannomutase algH Unknown function algR Regulatory component of two-component signal transduction system (cognate sensor kinase is FimS) algQ Histone-like transcriptional regulator binds to algD promoter algP Histone-like transcriptional regulator binds to algD promoter algZ AlgR cognate sensor (alginate and motility regulator) algU Homologous to Escherichia coli sE global stress response factor mucA Anti s factor mucB Anti s factor? mucC Regulator? mucD Homologous to E. coli serine protease DegP algW Homologous to E. coli serine protease DegS mucP Homologous to E. coli RseP protease involved in activation of AlgU via regulated intramembrane proteolysis cascade mucE Periplasmic or outer-membrane protein involved in activation of AlgU via regulated intramembrane proteolysis cascade mucR Alginate specific c-di-GMP synthesizing enzyme Modified according to Rehm (2002) c-di-GMP bis-(3¢-5¢)-cyclic dimeric guanosine monophosphate a Putative function indicated by ?
The alginate biosynthesis genes encoding enzymes catalysing the synthesis of the precursor GDP-mannuronic acid from fructose 6-phosphate have all been functionally assigned and characterized (Table 1, Fig. 1) (Rehm and Valla 1997).
2.2
GDP-Mannuronic Acid Biosynthesis
Alginate biosynthesis enzyme activities were initially assessed in the brown alga Fucus gardneri by Lin and Hassid (1966) and 10 years later in A. vinelandii (Pindar and Bucke 1975). The central metabolite fructose 6-phosphate derived
60
B.H.A. Rehm
from gluconeogenesis was found to be a precursor for the synthesis of GDPmannuronic acid. The alginate-specific and bifunctional enzyme phosphomannose isomerase (PMI)/GDP-mannose pyrophosphorylase (GMP), encoded by the algA gene, catalyses the initial and third steps of GDP-mannuronic acid synthesis starting from fructose 6-phosphate (Fig. 1) (May et al. 1994). The PMI-catalysed reaction channels the fructose 6-phosphate towards alginate biosynthesis via the formation of mannose 6-phosphate (Shinabarger et al. 1991). Mannose 6-phosphate is then converted to mannose 1-phosphate by the phosphomannomutase (AlgC) (Zielinski et al. 1991). The AlgC enzyme has been found to show additionally phosphoglucomutase activity and its role in rhamnolipid and lipopolysaccharide biosynthesis was demonstrated experimentally (Goldberg et al. 1993; Olvera et al. 1999; Ye et al. 1994). The GMP activity of AlgA then converts the mannose 1-phosphate to GDPmannose, with concomitant hydrolysis of GTP (Shinabarger et al. 1991). AlgA has been studied in more detail and amino acids required for the GMP enzyme activity were identified through site-directed mutagenesis of the algA gene (May et al. 1994). Replacement of Lys-175 by arginine, glutamine or glutamate resulted in a 470–3,200-fold increased Km for mannose 1-phosphate when compared with the wildtype enzyme. These results suggested a role of Lys-175 in the binding of the substrate mannose 1-phosphate. Replacement of Arg-19 by glutamine, histidine or leucine resulted in a fourfold to sevenfold increased Km for GTP when compared with the wild-type enzyme. Hence, Arg-19 might be involved in the binding of GTP. Limited proteolysis analysis showed that the C terminus is essential for PMI activity but not for GMP activity, suggesting that the bifunctional PMI/GMP protein is composed of two independent enzymatic domains (May et al. 1994). GDP-mannose dehydrogenase (GMD), encoded by the algD gene, is a key enzyme in the biosynthesis of alginate and catalyses the almost irreversible oxidation of GDP-mannose to GDP-mannuronic acid. This is the committed step in alginate biosynthesis and it was shown to represent the metabolic bottleneck in alginateoverproducing mucoid strains of P. aeruginosa (Tatnell et al. 1993, 1994). Thus, AlgD has been conceived as a potential target to inhibit alginate production and hence to combat P. aeruginosa infections. GMD inhibitors have been identified and GMD inhibition has been shown to increase the susceptibility of a mucoid strain of P. aeruginosa to tobramycin, which is widely used for the treatment of CF lung infections (Snook et al. 2003). Thus, GMD inhibitors can be conceived as potential drugs which impair alginate production and the formation of characteristic biofilms and thereby strongly aid the efficiency of antibiotics. The absence of enzymes corresponding to AlgD in humans makes these inhibitors very specific by presumably avoiding severe side effects. GMD belongs to a small family of NAD+-dependent four-electron-transfer dehydrogenases, which include UDP-glucose dehydrogenase (UGD), histidinol dehydrogenase and 3-hydroxy-3-methylglutaryl-CoA reductase. UGD and GMD have been proposed to share a similar reaction mechanism by using a single active site to catalyse the two-step conversion of an alcohol to the respective acid via a thiohemiacetal intermediate. The structure of UGD from Streptococcus pyogenes has been resolved and showed that the enzyme forms a dimer (Campbell et al. 2000). Biochemical characterization of the P. aeruginosa GMD revealed
Alginate Production: Precursor Biosynthesis
61
an allosteric and cooperative behaviour, which suggested that the enzyme at least forms an oligomer composed of six subunits (Naught et al. 2002; Roychoudhury et al. 1989). In a more recent study, the crystal structure of the P. aeruginosa GMD in complex with its cofactor NAD(H) and reaction product, GDP-mannuronic acid, was determined at a resolution of 1.55 Å (Snook et al. 2003). The crystal structure was used to shed light on the multistep reaction catalysed by GMD. The reaction comprises four steps: (1) oxidation of a hydroxyl to an aldehyde; (2) nucleophilic attack by a thiol to form the thiohemiacetal intermediate; (3) oxidation of the intermediate to a thioester; (4) release of the product by hydrolysis. On the basis of amino acid sequence alignments and a structural comparison with UGD, the cysteine at position 268 has been proposed as the active-site nucleophilic thiol (Campbell et al. 2000; Roychoudhury et al. 1992). The significance of this residue was confirmed by the sensitivity of GMD to thioreactive agents (Shankar et al. 1995). Accordingly, the crystal structure of GMD showed that the Cys-268 thiol group and the oxygens of the carboxylate group of the mannuronic acid could be within 2.4 Å. Thus, the thiol group as a potent nucleophile presumably serves as the active site required for formation of the proposed thiohemiacetal intermediate.
3
Alginate Polymerization and Secretion
The polymerization of exopolysaccharides usually requires the activity of membranebound glycosyltransferases which catalyse the transfer of an activated sugar moiety onto a receptor molecule while forming a glycosidic bond. The Wz-dependent capsular polysaccharide polymerization/secretion pathways in Escherichia coli have been extensively studied and serve as a model for the biosynthesis of various bacterial exopolysacharides (Whitfield 2006). Repeating sugar units are synthesized attached to a lipid carrier by glycosyltransferases at the cytoplasmic membrane. The lipid carrier linked repeating unit is then transferred across the cytoplasmic membrane and polymerization occurs at the periplasmic side of the membrane. The polymerized repeating units are subsequently secreted through specific pores in the outer membrane (Whitfield 2006). However, alginate represents a non-repeating exopolysaccharide and no alginate-related lipid carrier intermediate could be identified so far. Hence, alginate polymerization and transfer across the membrane might be based on a different molecular mechanism. The polymerization and secretion of alginate are barely understood and extensive research will be required to shed light on these molecular processes (Rehm and Valla 1997; Rehm 2005).
3.1
Alginate Polymerization
The polymerization of mannuronic acid residues to alginate could resemble the lipid carrier independent cellulose synthesis in bacteria such as Gluconacetobacter xylinus (Ross et al. 1991). The cellulose synthase comprises a multiprotein complex of 420
62
B.H.A. Rehm
kDa which resides in the cytoplasmic membrane and catalyses the processive polymerization of glucose by using UDP-glucose as a substrate. New findings related to bacterial cellulose biosynthesis were recently summarized by Valla et al. (2009). The protein Alg8, presumably encoding a glycosyltransferase, and Alg44 were predicted to be transmembrane proteins as well as subunits of the alginate polymerase (Oglesby et al. 2008; Remminghorst and Rehm 2006b, c). Recently, Alg8 was identified as a key membrane protein for the production of alginate and multiple copies of Alg8 resulted in significant overproduction of alginate (Remminghorst and Rehm 2006c). The same study showed for the first time the in vitro synthesis of alginate by using an envelope fraction of P. aeruginosa. Interestingly, separation of cytoplasmic membrane and outer membrane abolished alginate polymerization activity, which strongly suggested a coordinated polymerization and secretion as well as the requirement of a multiprotein complex spanning the cytoplasmic membrane, the periplasm and the outer membrane (Fig. 3) (Remminghorst and Rehm 2006c). Additionally, the isogenic alg8 deletion mutant did not secrete uronic acids, which were proposed to be derived from alginate lyase (AlgL) mediated degradation of misguided and unprotected alginate when found to be produced by isogenic algK, algX and algG (epimerase gene) deletion mutants, respectively. Interestingly, an AlgL gene deletion mutant of P. aeruginosa showed a swollen periplasm shortly after induction of alginate production, which suggested that AlgL does not only degrade misguided alginate but also contributes to the proposed protein scaffold (Fig. 3) (Jain and Ohman 2005). In contrast to these proposed scaffold-forming proteins (AlgK, AlgX, AlgG, AlgL), which were proposed to play a role in guidance
Alginate
Outer membrane
AlgE AlgL
Periplasm
AlgX
AlgK
MucD
RIP cascade
AlgG Cytoplasmic membrane Alg8
Alg44
MucR
c-di-GMP GDP-mannuronic acid
2 GTP Alginate gene regulation
Fig. 3 Model of the proposed multiprotein complex involved in alginate polymerization/export. c-di-GMP bis-(3¢–5¢)-cyclic dimeric guanosine monophosphate, RIP regulated intramembrane proteolysis, Alg alginate gene (for a description see Table 1)
Alginate Production: Precursor Biosynthesis
63
and protection of the nascent alginate chain, Alg8 seems to be directly involved in alginate polymerization. Hydrophobic cluster analysis of Alg8 indicated similarities to processive b-glycosyltransferases (e.g. cellulose or chitin synthase) (Saxena et al. 1995), which are transmembrane proteins that catalyse the formation of polysaccharide or oligosaccharide chains by transferring the sugar residue from an activated donor substrate to a growing acceptor molecule. These b-glycosyltransferases are related on the basis of their sequence pattern and especially with respect to the presence of conserved motifs and catalytic residues (Saxena et al. 2001). The catalysis is believed to involve a general base, which assists in the deprotonation process of the nucleophilic hydroxyl of the acceptor, and an oxocarbenium ion like transition state similar to that proposed for glycosidases (Ünligil et al. 2000). A threading model of Alg8 was developed on the basis of the crystal structure of SpsA, a glycosyltransferase involved in spore coat polysaccharide formation of Bacillus subtilis (Charnock and Davies 1999; Remminghorst and Rehm 2006c). In a recent study, a refined model for the membrane topology of Alg8 was developed using PhoA (alkaline phosphatase) and LacZ (b-glactosidase) reporter enzyme fusions, respectively (Oglesby et al. 2008; Remminghorst et al. 2009). Evidence for a large cytoplasmic loop containing the active domains and five transmembrane domains as well as for a short periplasmic loop was obtained. The presence of a cytoplasmic loop was further confirmed by successful production of only the cytoplasmic domain as soluble protein (Remminghorst et al. 2009). The C-terminal transmembrane domain of Alg8 was found to be essential for the in vivo polymerization reaction. Conserved amino acid residues of Alg8 were subjected to site-specific mutagenesis. The predicted active-site residues in the D133, D188-x-D190, L336xxR339W340 motif as well as in D295/D296 and K297 were found to be required for in vivo polymerization activity (Oglesby et al. 2008; Remminghorst et al. 2009). Alg44 was recently demonstrated to be a membrane protein with a periplasmic C terminus and was found to be essential for alginate polymerization (Merighi et al. 2007; Remminghorst and Rehm 2006b). A refined membrane topology model of Alg44 was obtained by constructing further PhoA fusions which showed a central transmembrane domain (amino acid residues 159–177) preceded by an N-terminal transmembrane domain (amino acid residues 1–30) presumably serving as a membrane anchor (Oglesby et al. 2008). The N-terminal cytoplasmic loop comprises the PilZ domain, which is required for binding of the secondary messenger bis-(3¢–5¢)-cyclic dimeric guanosine monophosphate (c-di-GMP) (Fig. 3). The PilZ domain in Alg44 is essential for alginate biosynthesis, which was recently verified experimentally by assessing the capability of modified and truncated Alg44 to restore alginate biosynthesis in an alg44-negative mutant of P. aeruginosa (Merighi et al. 2007). Thus, similar to bacterial cellulose synthesis, c-di-GMP has been verified as an important activator of bacterial alginate biosynthesis which adds an additional posttranslational regulatory layer to the already complex regulatory network which controls alginate gene expression. Interestingly, more than 33 enzymes (diguanylate cyclases, phosphodiesterases) have been found to be involved in the control of intracellular levels of c-di-GMP in P. aeruginosa (Kulasakara et al. 2006). It has been proposed that these enzymes contrib-
64
B.H.A. Rehm
ute to the formation of localized c-di-GMP pools mediating differential regulation of various responses, such as flagella biogenesis and exopolysaccharide biosynthesis. To identify the diguanylate cyclase which provides a c-di-GMP pool for Alg44 and hence for activation of alginate, a putatively membrane anchored diguanylate cyclase PA1727 (designated MucR) which also comprises a sensor domain was analysed in more detail. An isogenic knockout of this gene caused a 38-fold decrease in alginate production in P. aeruginosa, which suggested that this diguanylate cyclase catalyses c-di-GMP formation required for activation of alginate biosynthesis (Hay et al., 2009). The extended periplasmic C terminus of Alg44 shows similarities to membrane fusion proteins such as MexA, a membrane-bridging protein involved in the multidrug efflux system of P. aeruginosa (Akama et al. 2004). Hence, Alg44 could function as a subunit of a periplasmic protein scaffold and/or link Alg8 in the cytoplasmic membrane with the outer membrane protein AlgE (Remminghorst and Rehm 2006b). Experimental evidence for a function of the C terminus of Alg44 in colocalization and stabilization of a protein complex comprising Alg8 in the cytoplasmic membrane and AlgE in the outer membrane was obtained by detecting reduced levels of AlgE in an alg44-negative mutant (Oglesby et al. 2008). A C-terminal deletion of only 24 amino acid residues of Alg44 did not enable restoration of alginate production, further suggesting that Alg44 might act as a copolymerase essential for alginate polymerization. AlgK and AlgX are periplasmic proteins and the respective deletion mutants showed secretion of free uronic acids presumably owing to AlgL-catalysed degradation of misguided periplasmic alginate (Gutsche et al. 2006; Jain and Ohman, 1998; Jain and Ohman 2005). A pull-down experiment with Strep-tag-labelled AlgX produced in an isogenic algX-negative mutant of P. aeruginosa resulted in copurification of MucD, which suggested that AlgX specifically interacts with the serine protease MucD (Gutsche et al. 2006). However, no evidence for interaction with the other proposed subunits of the alginate polymerase/secretion complex could be obtained (Gutsche et al. 2006). Since knockout mutants of the mucD gene resulted in a mucoid phenotype, MucD was proposed as a negative regulator of alginate biosynthesis (Boucher et al. 1996; Wood and Ohman 2006; Yorgey et al. 2001). The negative regulator function was recently shown to require the serine protease activity (Wood and Ohman 2006). AlgX as a subunit of the alginate polymerase/secretion complex might sequester MucD in the periplasm and thereby interfere with its negative regulatory function. In a recent study, a regulated intramembrane proteolysis (RIP) cascade in the envelope of P. aeruginosa was proposed that ultimately leads to degradation of membraneanchored MucA. Degradation of MucA causes the release of sequestered alginatespecific sigma-factor AlgU, which then enables transcription of the alginate biosynthesis genes (Qiu et al. 2007) (see “Alginate gene regulation” by Ohman, this volume, for further details). In the absence of AlgX, MucD could interfere with RIP activation of alginate gene transcription by degrading proteins such as AlgW, MucP, MucE and MucB (Fig. 3). These proteins have been proposed to be involved in the RIP cascade (Qiu et al. 2007).
Alginate Production: Precursor Biosynthesis
65
The AlgK sequence showed an apparent signal peptide characteristic of a lipoprotein. AlgK-PhoA as well as AlgK-b-lactamase fusion proteins were found to show reporter enzyme activity, which suggested a periplasmic subcellular localization while they were probably anchored in the cytoplasmic membrane (Aarons et al. 1997; Jain and Ohman 1998). Overall the amino acid sequence of AlgK does not indicate a specific function of the protein in alginate polymerization/secretion. However, the AlgK sequence shows four Sel1-like repeats (SLR) which are characteristic of proteins forming modular architectures and which all seem to serve as adaptor proteins for the assembly of macromolecular complexes (Mittl and Schneider-Brachert 2007). Thus, AlgK might play a key role in the assembly of the alginate polymerization/secretion multiprotein complex. Recently, AlgK was subjected to a structural analysis in an attempt to verify the presence of the putative SLR motifs and provide insight into the function of AlgK. AlgK could be crystallized and preliminary X-ray data were obtained (Keiski et al. 2007). On the basis of density calculations, the authors estimated that four molecules of the protein are present in the asymmetric unit (Keiski et al. 2007). Size-exclusion chromatography results using purified AlgK provided evidence that AlgK forms a dimer in solution, suggesting that two dimers constitute the asymmetric unit in crystallized AlgK. AlgX, AlgK, Alg44, AlgG together with AlgL were proposed to form a protein scaffold in the periplasm surrounding the nascent alginate chain and guiding the polymer to AlgE in the outer membrane (Fig. 3) (Gutsche et al. 2006; Jain and Ohman 1998, 2005; Jain et al. 2003; Remminghorst and Rehm 2006a, b), while AlgL might also clear the periplasm from misguided alginate and AlgG introduces guluronic acid residues (Bakkevig et al. 2005). Interestingly, except for the AlgXMucD interaction, none of the proposed protein-protein interactions in the alginate polymerization/secretion multiprotein complex were confirmed experimentally.
3.2
Alginate Export
An early comparative analysis of outer-membrane protein profiles from various mucoid and non-mucoid P. aeruginosa strains indicated the presence of a 54-kDa protein exclusively found in mucoid strains (Grabert et al. 1990); hence, it was proposed that this outer-membrane protein could play a role in alginate biosynthesis. Shortly after the detection of the mucoid-specific 54-kDa outer-membrane protein, the algE gene in P. aeruginosa was identified as an essential gene for alginate production and which is also part of the alginate biosynthesis gene cluster (Chu et al. 1991). In the same study the algE gene product was identified as an insoluble presumably membrane bound protein with an apparent molecular mass of 54 kDa. Purification of the 54-kDa protein after solubilization with octylglucoside enabled N-terminal sequencing which confirmed that the 54-kDa protein corresponded to AlgE (Grabert et al. 1991). The fact that AlgE is strictly associated with the alginate-overproducing mucoid phenotype of P. aeruginosa triggered a study investigating the immunogenicity of
66
B.H.A. Rehm
AlgE as well as the use AlgE as an antigen for antibody capture in diagnostic applications (Rehm et al. 1994b). Purified and native AlgE was injected into rabbits and a strong antibody response 9 weeks after injection suggested a strong immunogenicity of AlgE (Rehm et al. 1994b). Hence, it was conceived that AlgE could be used for antibody detection in sera from CF patients. An antibody-capture ELISA with AlgE as an antigen was established and the analysis of 41 sera from CF patient showed a strong correlation with the infection status of the CF patients (Rehm et al. 1994b). None of the 23 control sera from healthy humans showed significant levels of anti-AlgE antibodies. Interestingly, CF patients who became infected only a few weeks before the blood samples were taken already showed a significant anti-AlgE antibody titre, suggesting the conversion to the mucoid form occurs early after infection with non-mucoid forms (Rehm et al. 1994b). Since CF patients are initially infected by non-mucoid P. aeruginosa and the mucoid form, which correlates with the establishment of a chronic infection, emerges in the CF lung during infection, the AlgEbased ELISA might represent an important differential diagnostic tool to assess the infection status. The infection status significantly informs the antibacterial treatment regimen, because mucoid P. aeruginosa shows a different and mainly increased antibiotic resistance, when compared with the non-mucoid variant. The strong AlgE immunogenicity also suggested an application of this mucoid-specific outermembrane protein as part of a multicomponent experimental vaccine which could mediate protection against colonization with mucoid forms of P. aeruginosa (Rehm et al. 1994b). AlgE is only represented as a minor protein with a low copy number in the outer membrane. To produce significant amounts of AlgE for functional analysis, algE was overexpressed in E. coli using a T7-promoter-based overexpression system. Recombinant and mature AlgE was localized to the outer membrane of E. coli by using immunoblotting analysis. This analysis suggested that the signal peptide was properly cleaved off during the secretion process, while traces of unprocessed AlgE were still detectable in the cytoplasmic membrane (Rehm et al. 1994a). Recombinant AlgE was solubilized from the E. coli outer membrane and purified by immobilized metal ion affinity chromatography followed by anion-exchange chromatography. Outer-membrane proteins have been described to function as general or specific porins, which form a b-barrel and facilitate the transport of various or specific compounds through the outer membrane (Buchanan 1999; Delcour 2002; Schulz 1996). Thus, to assess whether AlgE functions as such a channel protein it was subjected to electrophysiological analysis using planar lipid bilayers. These experiments demonstrated that AlgE can spontaneously incorporate into planar lipid bilayers while causing single-channel current fluctuations with a low mean conductance of 0.76 nS and a very short mean open-state lifetime of 0.7 ms (Rehm et al. 1994a). The electrophysiological analysis suggested that AlgE forms a channel protein in the outer membrane with unusual properties hitherto not described for other outer-membrane channel proteins (Delcour 2002). The AlgE channel was characterized in more detail by using differently sized cations and anions, respectively. Only the size of the anion had an impact on the mean single-channel conductance and increasing anion size lowered the mean single-channel conductance. These findings suggested that AlgE
Alginate Production: Precursor Biosynthesis
67
forms an anion-specific channel which could facilitate the export of the anionic alginate through the outer membrane. This proposed function of AlgE was further supported by the finding that GDP-mannuronic acid can partially block the AlgE channel (Rehm et al. 1994a). Topological models of AlgE and its homologue AlgJ from A. vinelandii have been developed on the basis of secondary structure predictions, hydrophilicity analysis and known outer membrane protein structures, resulting in the proposal of a b-barrel consisting of 18 b-strands (Fig. 3) (Rehm et al. 1994a; Rehm 1996). Recently, homology modelling was used to generate a three-dimensional model of AlgE suggesting a pore diameter suitable for alginate export (Rehm 2002). These data supported the hypothesis that AlgE forms an alginate-specific pore which enables export of the nascent alginate chain through the outer membrane. The AlgE protein is the only outer-membrane protein which has been found to be required for alginate biosynthesis and which might be colocalized with the alginate polymerase by the membrane fusion protein function of Alg44.
4
Conclusion and Future Perspectives
Although the alginate biosynthesis gene cluster was identified in the late 1980s, the functional assignment of various genes presumably contributing to alginate polymerization/secretion is still at an early stage. Except for the AlgX–MucD interaction, none of the proposed protein-protein interactions required to form an envelope-spanning multiprotein complex have been confirmed experimentally. Further extensive research will be required to shed light on the molecular mechanisms underlying alginate polymerization and secretion as well as the c-di-GMP-mediated activation. The understanding of the final steps in alginate production might inform the design of specific inhibitors when considering the role of alginate as a virulence factor and/or it might help to overcome the production bottleneck when considering alginate as an important biopolymer for medical and biotechnological applications. Although commercial production of alginate relies almost entirely on cheap algal sources, the ability to engineer bacterial alginates will make continuous fermentative production using bacteria increasingly attractive. Knowledge of how alginate composition and molecular weight contribute to material properties and knowledge of what material properties are required to meet medical specifications will provide an enormous opportunity for the use of engineered bacteria for the production of tailor-made alginates.
References Aarons SJ, Sutherland IW, Chakrabarty AM, Gallagher MP (1997) A novel gene, algK, from the alginate biosynthesis cluster of Pseudomonas aeruginosa. Microbiology 143:641–652 Akama H, Matsuura T, Kashiwagi S, Yoneyama H, Narita S, Tsukihara T, Nakagawa A, Nakae T (2004) Crystal structure of the membrane fusion protein, MexA, of the multidrug transporter in Pseudomonas aeruginosa. J Biol Chem 279:25939–25942
68
B.H.A. Rehm
Bakkevig K, Sletta H, Gimmestad M, Aune R, Ertesvag H, Degnes K, Christensen BE, Ellingsen TE, Valla S (2005) Role of the Pseudomonas fluorescens alginate lyase (AlgL) in clearing the periplasm of alginates not exported to the extracellular environment. J Bacteriol 187: 8375–8384 Boucher JC, Martinez-Salazar J, Schurr MJ, Mudd MH, Yu H, Deretic V (1996) Two distinct loci affecting conversion to mucoidy in Pseudomonas aeruginosa in cystic fibrosis encode homologs of the serine protease HtrA. J Bacteriol 178:511–523 Buchanan SK (1999) Beta-barrel proteins from bacterial outer membranes: structure, function and refolding. Curr Opin Struct Biol 9:455–461 Campbell RE, Mosimann SC, van De Rijn I, Tanner ME, Strynadka NC (2000) The first structure of UDP-glucose dehydrogenase reveals the catalytic residues necessary for the two-fold oxidation. Biochemistry 39:7012–7023 Charnock SJ, Davies GJ (1999) Structure of the nucleotide-diphospho-sugar transferase, SpsA from Bacillus subtilis, in native and nucleotide-complexed forms. Biochemistry 38:6380–6385 Chitnis CE, Ohman DE (1993) Genetic analysis of the alginate biosynthetic gene cluster of Pseudomonas aeruginosa shows evidence of an operonic structure. Mol Microbiol 8:583–593 Chu L, May TB, Chakrabarty AM, Misra TK (1991) Nucleotide sequence and expression of the algE gene involved in alginate biosynthesis by Pseudomonas aeruginosa. Gene 107:1–10 Darzins A, Chakrabarty AM (1984) Cloning of genes controlling alginate biosynthesis from a mucoid cystic fibrosis isolate of Pseudomonas aeruginosa. J Bacteriol 159:9–18 Delcour AH (2002) Structure and function of pore-forming beta-barrels from bacteria. J Mol Microbiol Biotechnol 4:1–10 Gacesa P (1998) Bacterial alginate biosynthesis – recent progress and future prospects. Microbiology 144:1133–1143 Gaona G, Nunez C, Goldberg JB, Linford AS, Najera R, Castaneda M, Guzman J, Espin G, Soberon-Chavez G (2004) Characterization of the Azotobacter vinelandii algC gene involved in alginate and lipopolysaccharide production. FEMS Microbiol Lett 238:199–206 Goldberg JB, Hatano K, Pier GB (1993) Synthesis of lipopolysaccharide O side chains by Pseudomonas aeruginosa PAO1 requires the enzyme phosphomanno-mutase. J Bacteriol 175:1605–1611 Grabert E, Wingender J, Winkler UK (1990) An outer membrane protein characteristic of mucoid strains of Pseudomonas aeruginosa. FEMS Microbiol Lett 56:83–87 Grabert E, Boheim G, Winkler UK (1991) Biochemical and functional properties of mucoidcharacteristic 54 kDa outer membrane protein of Pseudomonas aeruginosa. In: Pseudomonas: third international symposium on pseudomonads, biology and biotechnology (abstracts), Miramare-Griguano, Trieste, Italy, p 99 Gutsche J, Remminghorst U, Rehm BHA (2006) Biochemical analysis of alginate biosynthesis protein AlgX from Pseudomonas aeruginosa: purification of an AlgX-MucD (AlgY) protein complex. Biochimie 88:245–251 Hay ID, Remminghorst U, Rehm BHA (2009) MucR, a novel membrane associated regulator of alginate biosynthesis in Pseudomonas aeruginosa. Appl Environ Microbiol (in press). Jain S, Ohman DE (1998) Deletion of algK in mucoid Pseudomonas aeruginosa blocks alginate polymer formation and results in uronic acid secretion. J Bacteriol 180:634–641 Jain S, Franklin MJ, Ertesvag H, Valla S, Ohman DE (2003) The dual roles of AlgG in C-5epimerization and secretion of alginate polymers in Pseudomonas aeruginosa. Mol Microbiol 47:1123–1133 Jain S, Ohman DE (2005) Role of an alginate lyase for alginate transport in mucoid Pseudomonas aeruginosa. Infect Immun 73:6429–6436 Keiski CL, Yip P, Robinson H, Burrows LL, Howell PL (2007) Expression, purification, crystallization and preliminary X-ray analysis of Pseudomonas fluorescens AlgK. Acta Crystallogr 63:415–418 Kobayashi H (2005) Airway biofilms: implications for pathogenesis and therapy of respiratory tract infections. Treat Respir Med 4:241–253 Kulasakara H, Lee V, Brencic A, Liberati N, Urbach J, Miyata S, Lee DG, Neely AN, Hyodo M, Hayakawa Y, Ausubel FM, Lory S (2006) Analysis of Pseudomonas aeruginosa diguanylate
Alginate Production: Precursor Biosynthesis
69
cyclases and phosphodiesterases reveals a role for bis-(3¢-5¢)-cyclic-GMP in virulence. Proc Natl Acad Sci U S A 103:2839–2844 Ledgham F, Soscia C, Chakrabarty A, Lazdunski A, Foglino M (2003) Global regulation in Pseudomonas aeruginosa: the regulatory protein AlgR2 (AlgQ) acts as a modulator of quorum sensing. Res Microbiol 154:207–213 Lin TY, Hassid WZ (1966) Pathway of alginic acid synthesis in the marine brown alga, Fucus gardneri Silva. J Biol Chem 241:5284–5297 Lizewski SE, Schurr JR, Jackson DW, Frisk A, Carterson AJ, Schurr MJ (2004) Identification of AlgR-regulated genes in Pseudomonas aeruginosa by use of microarray analysis. J Bacteriol 186:5672–5684 May TB, Shinabarger D, Maharaj R, Kato J, Chu L, Devault JD, Roychoudhury S, Zielinsky NA, Berry A, Rothmel RK, Misra TK, Chakrabarty AM (1991) Alginate synthesis by Pseudomonas aeruginosa: a key pathogenic factor in chronic pulmonary infections of cystic fibrosis patients. Clin Microbiol Rev 4:191–206 May TB, Shinabarger D, Boyd A, Chakrabarty AM (1994) Identification of amino acid residues involved in the activity of phosphomannose isomerase-guanosine 5¢-diphospho-d-mannose pyrophosphorylase. A bifunctional enzyme in the alginate biosynthetic pathway of Pseudomonas aeruginosa. J Biol Chem 269:4872–4877 Merighi M, Lee VT, Hyodo M, Hayakawa Y, Lory S (2007) The second messenger bis-(3¢-5¢)cyclic-GMP and its PilZ domain-containing receptor Alg44 are required for alginate biosynthesis in Pseudomonas aeruginosa. Mol Microbiol 65:876–895 Mittl PR, Schneider-Brachert W (2007) Sel1-like repeat proteins in signal transduction. Cell Signal 19:20–31 Naught LE, Gilbert S, Imhoff R, Snook C, Beamer L, Tipton P (2002) Allosterism and cooperativity in Pseudomonas aeruginosa GDP-mannose dehydrogenase. Biochemistry 41:9637–9645 Oglesby LL, Jain S, Ohman DE (2008) Membrane topology and roles of Pseudomonas aeruginosa Alg8 and Alg44 in alginate polymerization. Microbiology 154:1605–1615 Olvera C, Goldberg JB, Sanchez R, Soberon-Chavez G (1999) The Pseudomonas aeruginosa algC gene product participates in rhamnolipid biosynthesis. FEMS Microbiol Lett 179:85–90 Paul F, Morin A, Monsan P (1986) Microbial polysaccharides with actual potential industrial applications. Biotechnol Adv 4:245–259 Pindar D, Bucke C (1975) Biosynthesis of alginic acid bu Azotobacter vinelandii. Biochem J 152:617–622 Qiu D, Eisinger VM, Rowen DW, Yu HD (2007) Regulated proteolysis controls mucoid conversion in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 104:8107–8112 Rehm BHA (1996) The Azotobacter vinelandii gene algJ encodes an outer-membrane protein presumably involved in export of alginate. Microbiology 142:873–880 Rehm BHA (2002) Alginates from bacteria. Biopolymers 8:179–212 Rehm BHA (2005a) Alginates from bacteria. In: Steinbüchel A, Marchessault RE (eds) Biopolymers for medical and pharmaceutical applications, vol 1. Wiley, Chichester, pp 265–298 Rehm BHA (2005b) Biosynthesis and applications of alginates. Biosynthesis and applications of alginates. In: Wnek G, Bowlin G (eds) Encyclopedia of biomaterials and biomedical engineering. Dekker, New York, pp 1–9 Rehm BHA (2009) Pseudomonas applications. Encyclopedia of Industrial Biotechnology, in press Rehm BHA, Valla S (1997) Bacterial alginates: biosynthesis and applications. Appl Microbiol Biotechnol 48:281–288 Rehm BHA, Boheim G, Tommassen J, Winkler UK (1994a) Overexpression of algE in Escherichia coli: subcellular localization, purification, and ion channel properties. J Bacteriol 176:5639–5647 Rehm BHA, Grabert E, Hein J, Winkler UK (1994b) Antibody response of rabbits and cystic fibrosis patients to an alginate-specific outer membrane protein of a mucoid strain of Pseudomonas aeruginosa. Microb Pathog 16:43–51 Remminghorst U, Rehm BHA (2006a) Bacterial alginates: from biosynthesis to applications. Biotechnol Lett 28:1701–1712
70
B.H.A. Rehm
Remminghorst U, Rehm BHA (2006b) Alg44, a unique protein required for alginate biosynthesis in Pseudomonas aeruginosa. FEBS Lett 580:3883–3888 Remminghorst U, Rehm BHA (2006c) In vitro alginate polymerization and the functional role of Alg8 in alginate production by Pseudomonas aeruginosa. Appl Environ Microbiol 72:298–305 Remminghorst U, Hay ID, Rehm BHA (2009) Molecular characterization of Alg8, a putative glycosyltransferase, involved in alginate polymerization. J Biotechnol (in press) Ross P, Mayer R, Benziman M (1991) Cellulose biosynthesis and function in bacteria. Microbiol Rev 55:35–58 Roychoudhury S, May TB, Gill JF, Singh SK, Feingold DS, Chakrabarty AM (1989) Purification and characterization of guanosine diphospho-d-mannose dehydrogenase. A key enzyme in the biosynthesis of alginate by Pseudomonas aeruginosa. J Biol Chem 264:9380–9385 Roychoudhury S, Chakrabarty K, Ho YK, Chakrabarty AM (1992) Characterization of guanosine diphospho-d-mannose dehydrogenase from Pseudomonas aeruginosa. Structural analysis by limited proteolysis. J Biol Chem 267:990–996 Russell NJ, Gacesa P (1988) Chemistry and biology of the alginate of mucoid strains of Pseudomonas aeruginosa in cystic fibrosis. Mol Aspects Med 10:1–91 Saxena IM, Brown RM Jr, Fevre M, Geremia RA, Henrissat B (1995) Multidomain architecture of beta-glycosyl transferases: implications for mechanism of action. J Bacteriol 177:1419–1424 Saxena IM, Brown RM Jr, Dandekar T (2001) Structure–function characterization of cellulose synthase: relationship to other glycosyltransferases. Phytochemistry 57:1135–1148 Schulz GE (1996) Porins: general to specific, native to engineered passive pores. Curr Opin Struct Biol 6:485–490 Shankar S, Ye RW, Schlictman D, Chakrabarty AM (1995) Exopolysaccharide alginate synthesis in Pseudomonas aeruginosa: enzymology and regulation of gene expression. Adv Enzymol Relat Area Mol Biol 70:221–255 Shinabarger D, Berry A, May TB, Rothmel R, Fialho A, Chakrabarty AM (1991) Purification and characterization of phosphomannose isomerase-guanosine diphospho-d-mannose pyrophosphorylase. A bifunctional enzyme in the alginate biosynthetic pathway of Pseudomonas aeruginosa. J Biol Chem 266:2080–2088 Skjak-Braek G, Grasdalen H, Larsen B (1986) Monomer sequence and acetylation pattern in some bacterial alginates. Carbohydr Res 154:239–250 Snook CF, Tipton PA, Beamer LJ (2003) Crystal structure of GDP-mannose dehydrogenase: a key enzyme of alginate biosynthesis in P. aeruginosa. Biochemistry 42:4658–4668 Stover CK, Pham XQ, Erwin AL, Mizoguchi SD, Warrener P, Hickey MJ, Brinkman FS, Hufnagle WO, Kowalik DJ, Lagrou M, Garber RL, Goltry L, Tolentino E, Westbrock-Wadman S, Yuan Y, Brody LL, Coulter SN, Folger KR, Kas A, Larbig K, Lim R, Smith K, Spencer D, Wong GK, Wu Z, Paulsen IT, Reizer J, Saier MH, Hancock REW, Lory S, Olson MV (2000) Complete genome sequence of Pseudomonas aeruginosa PA01, an opportunistic pathogen. Nature 406:959–964 Tatnell PJ, Russell NJ, Gacesa P (1993) A metabolic study of the activity of GDP-mannose dehydrogenase and concentrations of activated intermediates of alginate biosynthesis in Pseudomonas aeruginosa. J Gen Microbiol 139:119–127 Tatnell PJ, Russell NJ, Gacesa P (1994) GDP-mannose dehydrogenase is the key regulatory enzyme in alginate biosynthesis in Pseudomonas aeruginosa: evidence from metabolite studies. Microbiology 140:1745–1754 Ünligil UM, Zhou S, Yuwaraj S, Sarkar M, Schachter H, Rini JM (2000) X-ray crystal structure of rabbit N-acetylglucosaminyltransferase I: catalytic mechanism and a new protein superfamily. EMBO J 19:5269–5280 Valla S, Ertesvag H, Tonouchi N, Fjaervig E (2009) Bacterial cellulose production: Biosynthesis and applications. In: Rehm BHA (ed) Microbial production of biopolymers and polymer precursors: applications and perspectives. Horizon Bioscience, Caister Academic, Wymondham Whitfield C (2006) Biosynthesis and assembly of capsular polysaccharides in Escherichia coli. Annu Rev Biochem 75:39–68
Alginate Production: Precursor Biosynthesis
71
Wood LF, Ohman DE (2006) Independent regulation of MucD, an HtrA-like protease in Pseudomonas aeruginosa, and the role of its proteolytic motif in alginate gene regulation. J Bacteriol 188: 3134–3137 Wu W, Badrane H, Arora S, Baker HV, Jin S (2004) MucA-mediated coordination of type III secretion and alginate synthesis in Pseudomonas aeruginosa. J Bacteriol 186:7575–7585 Ye RW, Zielinski NA, Chakrabarty AM (1994) Purification and characterization of phosphomannomutase/phosphoglucomutase from Pseudomonas aeruginosa involved in biosynthesis of both alginate and lipopolysaccharide. J Bacteriol 176:4851–4857 Yorgey P, Rahme LG, Tan MW, Ausubel FM (2001) The roles of mucD and alginate in the virulence of Pseudomonas aeruginosa in plants, nematodes and mice. Mol Microbiol 41:1063–1076 Zielinski NA, Chakrabarty AM, Berry A (1991) Characterization and regulation of the Pseudomonas aeruginosa algC gene encoding phosphomannomutase. J Biol Chem 266:9754–9763
Bacterial System for Alginate Uptake and Degradation Wataru Hashimoto, Yukie Maruyama, Takafumi Itoh, Bunzo Mikami, and Kousaku Murata
Contents 1 2
Introduction.......................................................................................................................... Alginate-Assimilating Bacterium ........................................................................................ 2.1 Pit Formation on the Cell Surface ............................................................................... 2.2 Genome Sequence....................................................................................................... 3 Alginate Uptake ................................................................................................................... 3.1 Cell Surface Receptors................................................................................................ 3.2 Outer-Membrane Transporters .................................................................................... 3.3 Periplasmic Binding Proteins...................................................................................... 3.4 Inner-Membrane ABC Importer ................................................................................. 4 Alginate Degradation ........................................................................................................... 4.1 Endotype Alginate Lyases .......................................................................................... 4.2 Exotype Alginate Lyases ............................................................................................ 5 Concluding Remarks and Future Perspectives .................................................................... 5.1 Superchannel as Alginate Uptake System .................................................................. 5.2 Molecular Transplantation of Superchannel ............................................................... 5.3 Application of the Alginate Degradation System ....................................................... References ..................................................................................................................................
74 75 75 77 77 79 80 81 83 84 84 86 87 87 88 90 91
Abstract Most alginate-assimilating bacteria secrete degrading enzymes, i.e., alginate lyases, into the extracellular fraction or periplasm, and incorporate the resultant alginate oligosaccharides through their cytoplasmic membrane. The Gram-negative bacterium Sphingomonas sp. A1 can directly incorporate the polysaccharide into the cytoplasm, without degradation, through a mouthlike pit on the cell surface, periplasmic binding proteins, and an ATP-binding cassette importer in the cytoplasmic W. Hashimoto, Y. Maruyama, T. Itoh and K. Murata( ) Laboratory of Basic and Applied Molecular Biotechnology, Graduate School of Agriculture, Kyoto University, Uji, Kyoto 611-0011, Japan e-mail: [email protected] B. Mikami Laboratory of Applied Structural Biology, Graduate School of Agriculture, Kyoto University, Uji, Kyoto, 611-0011, Japan
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_3, © Springer-Verlag Berlin Heidelberg 2009
73
74
W. Hashimoto et al.
membrane. This uptake system (superchannel) is distinct from channels or transporters responsible for importing low molecular weight substrates. The constituent proteins are inducibly expressed and organized at the superchannel when flagellin homologues, as cell surface receptors, recognize the external alginate. Cytoplasmic alginate lyases with different substrate specificities and action modes help degrade the alginate into monosaccharides. The strain A1 superchannel can be transplanted to other sphingomonads through membrane engineering. This chapter reviews the bacterial system for alginate uptake and degradation by considering the structure and function of each molecule in the superchannel.
1
Introduction
Alginate is a linear polysaccharide produced by brown seaweed and certain bacteria. Two monosaccharides, b-d-mannuronate (M) and its C-5 epimer a-l-guluronate (G), constitute the polysaccharide molecules, and are arranged in three regions: polyM, polyG, and heteropolymeric random sequences (polyMG) (Gacesa 1988). Brown seaweed alginate is widely used in the food, chemical, and pharmaceutical industries as a stabilizing, thickening, or emulsifying reagent because the polymer chelates metal ions and forms a highly viscous solution (Jensen 1993). Oligosaccharides derived from brown seaweed alginate function physiologically as growth factors for microbial, plant, and animal cells, such as bifidobacteria, brassica, and human keratinocyte (Akiyama et al. 1992; Kawada et al. 1997, 1999; Iwamoto et al. 2003, 2005). Cytokine is also inducibly produced from mouse macrophage cells in the presence of the oligosaccharides (Kurachi et al. 2005). Two Gram-negative bacterial species, Azotobacter vinelandii and Pseudomonas aeruginosa, produce and secrete alginate extracellularly. Distinct from the seaweed polymer, bacterial alginate molecules are often acetylated at the mannuronate residue (Remminghorst and Rehm 2006). During the conversion of A. vinelandii vegetative cells to desiccation-resistant cysts, the extracellular alginate becomes a major component of the cyst exine (Moreno et al. 1998). An opportunistic pathogen, P. aeruginosa, produces extracellular biofilms, including alginate, as important virulence factors during lung infections in cystic fibrosis patients (May and Chakrabarty 1994). These biofilms often protect P. aeruginosa cells from phagocytic cells and/or antibiotics, making biofilm-dependent diseases difficult to treat (Schweizer and Boring 1973). Alginate-depolymerizing enzymes thus hold promise as biochemicals for processing edible seaweed alginate and removing bacterial biofilm alginate. A large number of alginate-assimilating microbes have been isolated from the soil, seawater, or wastewater, and some marine algae and mollusks have been found to produce alginate-degrading enzymes (Wong et al. 2000). To the best of our knowledge, all of the alginate-degrading enzymes that have been characterized to date strictly catalyze a b-elimination reaction, not a hydrolytic reaction, thus indicating that the enzymes are exclusively categorized as lyases. Although both polysaccharide lyases and hydrolases cleave the glycosidic bond in substrate molecules, the products released
Bacterial System for Alginate Uptake and Degradation
75
through the lyase reactions have unsaturated saccharides at the nonreducing terminus, with a double bond between C-4 and C-5. There are various types of substrate specificity, i.e., a preference for polyM, polyG, and/or polyMG. There are also two modes of action among the alginate lyases, endo or exo. When microbes utilize macromolecules for their growth, such as polysaccharides and proteins, the microbial cells commonly excrete degrading enzymes across the cytoplasmic membrane and then incorporate the degraded low molecular weight products into the cytoplasm. This is true of most alginate-assimilating microbes, and no microbial direct uptake system for alginate without degradation has ever been demonstrated, except for Sphingomonas sp. A1 (strain A1). Strain A1 has an unexplored specific and elegant system for alginate assimilation (Momma et al. 2000). Strain A1 also produces several alginate lyases with different substrate specificities and modes of action in the cytoplasm. The structure and function relationships of alginate lyases have been clarified through X-ray crystallography (Yoon et al. 1999; Yamasaki et al. 2005). This chapter primarily deals with the structural and functional bases for the strain A1 system for the direct uptake and degradation of alginate. It also discusses applications of the bacterial system to bioremediation, sugar technology, and biofuel generation.
2
Alginate-Assimilating Bacterium
Strain A1 belongs to the genus Sphingomonas, reclassified as a new member of the Proteobacteria a-4 subclass from Pseudomonas in 1990. Sphingomonads, bacteria in the genus Sphingomonas, including strain A1, are Gram-negative rods, aerobic, nonfermentative, polymixin B-resistant, and yellow-pigmented (White et al. 1996). Strain A1 is the only sphingomonad that assimilates alginate. Recently, phylogenetic, biochemical, and taxonomical analyses have been used to divide sphingomonads into four different genera, i.e., Novosphingobium, Sphingobium, Sphingomonas sensu stricto, and Sphingopyxis (Takeuchi et al. 2001).
2.1
Pit Formation on the Cell Surface
Sphingomonads are peculiar among Gram-negative bacteria in that their outer membrane includes glycosphingolipids (GSL) with long-chain base dihydrosphingosin in place of lipopolysaccharides (Kawasaki et al. 1994). In fact, some sphingomonads have been specifically shown to have many large pleatlike substances on their cell surface (Hashimoto et al. 2005c). Possibly owing to the presence of GSL-dependent pleatlike substances, specific cell surface properties for sphingomonads have been reported (White et al. 1996; Kinjo et al. 2005), e.g., hydrophobicity and antigenicity. In addition to the presence of GSL in the outer membrane, strain A1 has two
76
W. Hashimoto et al.
characteristic features, pit formation on the cell surface and no secretion of alginate lyases from the cytoplasm. Electron microscopy indicates that strain A1 also has many large pleats on the cell surface (Fig. 1a) (Hisano et al. 1996). Distinct from those of other sphingomonads, strain A1 cells grown on alginate form a mouthlike pit on the cell surface, with a size
a
b
0.01µm
c
0.01µm
0.01µm
d
0.01µm
Fig. 1 Electron micrographs of Sphingomonas sp. A1. a Strain A1 cell grown in the absence of alginate; b strain A1 cell grown on alginate; c p6-disruptant cell grown on alginate; d localization of flagellin homologue p5 in strain A1 cells by immunogold electron microscopy (left strain A1 cell grown on alginate; right strain A1 cell grown in the absence of alginate)
Bacterial System for Alginate Uptake and Degradation
77
of around 0.1 mm (Fig. 1b). This cell surface pit is formed through the reorganization and/or fluidity of the pleats. As far as we know, this is the first finding of a pit-forming bacterium in the history of microbiology. The pit formation is dependent on the presence of alginate in the external milieu, i.e., the pit forms in the presence of alginate, but disappears in the absence of the polymer. That is, the pit formation is reversible. Electron microscopy with alginate-staining dye has shown alginate concentrated in the pit, indicating that the pit functions as a funnel or concentrator for alginate. Since strain A1 secretes no alginate lyases, these bacterial cells must be equipped with a direct uptake system for alginate without degradation.
2.2
Genome Sequence
To clarify the direct uptake system for alginate in strain A1, the bacterial genome sequence was completely determined (Hashimoto et al. 2005c). This is the first sphingomonad to have its genome sequence determined. Strain A1 also has a circular double-stranded DNA plasmid (pA1) other than a chromosome (Harada et al. 2006). The chromosome of strain A1 comprises a circular double-stranded DNA consisting of 4,622,788 base pairs encoding about 4,800 genes. The GC content is relatively high (62%) among Gram-negative bacteria. The significant homology observed in 30% of whole genes between strain A1 and P. aeruginosa (Stover et al. 2000) is probably due to the reclassification of sphingomonads from pseudemonads, suggesting that strain A1 has evolved from pseudomonads. A large number of genes coding for alginate and pectin-degrading enzymes are found in the chromosome of strain A1. Similar to alginate, pectin is also a polyuronate. Strain A1 has also been experimentally demonstrated to assimilate pectin by cytoplasmic pectin-degrading enzymes, suggesting that strain A1 has become somewhat specialized to incorporate polysaccharides in its growth and has developed a direct uptake system. In fact, the genes responsible for alginate uptake and degradation are assembled into a cluster, as described later (Fig. 2) The plasmid pA1 comprises 46,557 base pairs coding for 49 genes with a GC content of 65% (Harada et al. 2006). The genes encoded in pA1 show a significant sequence similarity to those in the self-transmissible promiscuous incompatibility (Inc) group P-1b plasmid. The genetic organization of pA1 is also comparable to that of the IncP-1b plasmid. However, unlike any reported IncP-1 plasmids, pA1 contains no inserted mobile genetic elements.
3
Alginate Uptake
The cell surface pit is responsible for the direct uptake of alginate in strain A1. The formation and disappearance of the pit on the bacterial cell surface are, respectively, dependent on the presence and absence of alginate in the external milieu. This indicates
78
W. Hashimoto et al.
Fig. 2 Superchannel system for alginate uptake and degradation in strain A1. G l-guluronate, M d-mannuronate, aly gene for alginate lyases (A1-I, A1-II, and A1-III), ccpA catabolite-control protein gene, algS, algM1, and algM2 ABC importer genes for alginate import, algQ1 and algQ2 genes for alginate-binding proteins, a1-IV alginate lyase A1-IV gene, a1-II¢ alginate lyase A1-II¢ gene, a1-IV¢ alginate lyase A1-IV¢ gene, Algp3 TonB-dependent transporter, Algp5 alginate receptor
that comparative analyses, using transcriptomics and proteomics, of both bacterial cells grown in the presence of alginate and bacterial cells grown in its absence would contribute to a clarification of the direct uptake system for alginate.
Bacterial System for Alginate Uptake and Degradation
3.1
79
Cell Surface Receptors
Two flagellin homologues, p5 (40 kDa) and p6 (31 kDa), have been found to be inducibly expressed in the outer membrane by two-dimensional polyacrylamide gel electrophoresis dependent proteome analysis (Hashimoto et al. 2005a). Although flagellin is well known to be a component of the helical filament of flagella, strain A1 shows neither cell motility nor flagella formation. A p6 gene disruptant shows a significant growth retardation in a medium using alginate as the sole carbon source, although the disruption of the p5 gene has no effect on the growth of strain A1, even in the presence of alginate. The double disruption of the p5 and p6 genes results in growth failure, indicating that these flagellin homologues might be essential for cell viability. There are significant cell surface structural differences between the p6 gene disruptant and the wild type of strain A1. The gene disruptant showed incomplete pit formation and a network structure on the cell surface rather than a pleated one (Fig. 1c), suggesting that flagellin homologues regulate cell surface structures in strain A1. Thus, strain A1 flagellin homologues were overexpressed in Escherichia coli cells, purified, and characterized (Hashimoto et al. 2005a). Immunoelectron microscopy using anti-p5 antibody indicated that p5 was inducibly expressed in alginate-grown strain A1 cells and exclusively localized in the cell envelope (Fig. 1d, left). Strain A1 cells grown in the absence of alginate produced only small amounts of flagellin homologues (Fig. 1d, right). Since p5 was expressed as a cell surface protein in the presence of alginate, the interaction between p5 and alginate was analyzed using a surface plasmon resonance biosensor. Both proteins, p5 and p6, specifically bound to alginate with high affinity and two alginate-binding sites were located in the flagellin homologues. The dissociation constants ( Kd) in the two sites for p5 and alginate were determined to be 1.3 × 10−7 and 2.6 × 10−9 M. The high affinity of p5 for alginate may suggest that the protein functions as a cell surface receptor for alginate as is seen in a CD44-like receptor, a mammalian transmembrane receptor for hyaluronan (Turley et al. 2002). To clarify the structure and function relationship of the cell surface receptor for alginate, the crystal structure of truncated p5 (p5DN53C45) was determined at 2.0 Å resolution (Maruyama et al. 2008) (Fig. 2). The truncated p5 (p5DN53C45) had mutations with a lack of N-terminal 53 residues and C-terminal 45 residues. Two regions (flagellin_N and flagellin_C motifs) at the N- and C-termini were conserved in bacterial flagellins, while the central domains (flagellin_IN motif) varied to some extent. The flagellins were, therefore, divided into two groups, on the basis of the presence or absence of the flagellin_IN motif. p5DN53C45 is the first structure of a flagellin_IN-motif-containing flagellin to be determined, although the crystal structure of the Salmonella typhimurium flagellin, which lacks the flagellin_IN motif, was previously determined (Yonekura et al. 2003). p5DN53C45 consists of an a-domain rich in a-helices and a b-domain rich in b-strands. The a-domain is composed of N- and C-terminal regions, while the b-domain has a central region. The a-domain is structurally similar to the D1 domain of the S. typhimurium flagellin. The b-domain has a structure that resembles the finger domain of the
80
W. Hashimoto et al.
bacteriophage T4 baseplate protein (Leiman et al. 2004). The deletion mutant analysis suggested that residues 20–40 and 353–363 are involved in alginate binding. Since p5DN53C45 includes no alginate-binding regions, a structural analysis of the cell surface receptor with a whole sequence is now in progress. In addition to p5, the E. coli flagellin (FliC) also interacted with alginate, suggesting that alginate binding is clearly common to the bacterial flagellins (Hashimoto et al. 2005a). Bacterial flagellins are generally the flagella components responsible for cell motility (Aizawa 1996), and some pathogenic bacteria interact with hosts through flagellins (Hayashi et al. 2001). To the best of our knowledge, the behavior of the strain A1 flagellin homologues p5 and p6 is the first finding showing that a flagellin has physiological roles, especially alginate binding, in addition to its function as a propeller. This yields valuable hints to the origin, evolution, and function of flagella. One possibility is that a flagellin may be created first as a cell surface protein able to recognize nutrients present in the external milieu and/or to regulate cell surface structures, before changing into flagellar proteins, or vice versa. We can at least state that microbes with flagella can access materials containing alginate. But why alginate? Alginate is a polyanion and as such is able to interact with cations (Llanes et al. 2000). If, as we understand, alginate was the first polysaccharide produced in the sea early in biological history, and ferrous ions were abundant in the prebiotic ocean before oxygen was produced by cyanobacteria (Walker 1987), our results suggest the hypothesis that flagella were first created by ancient microbes to obtain alginate as a carbon and energy source for growth and/ or as a carrier for ferrous and other metal ions.
3.2
Outer-Membrane Transporters
Two proteins, p7 (28 kDa) and p8 (20 kDa), were overproduced in the outer membrane of alginate-grown strain A1 cells (He et al. 2004, 2008). Although p7 is similar to bacterial lipoproteins in its primary structure, it was found to have no lipid moiety by N-terminal sequence analysis and mass spectrometry. p7 was found to bind to alginate most efficiently at neutral pH with a Kd of 3.6 × 10−8 M. p8 is primarily similar to the polyhydroxyalkanoate granule associated protein of Ralstonia eutropha (Hanley et al. 1999). The disruptant of the p8 gene showed a significant growth retardation in an alginate medium. p8 was found to bind to alginate most efficiently at pH 4.0 with a Kd of 1.3 × 10−7 M. The p7 or p8 binding affinity for alginate was not much higher than that of the cell surface alginate receptor, suggesting that both proteins can associate and release the polymer. In addition to alginate molecules, p8 also exhibited an affinity for granules consisting of alginate and metal ions. The above-mentioned results indicated that p7 and p8 are outer-membrane proteins binding to alginate and facilitate the concentration of alginate in the pit on the cell surface of strain A1. Four proteins with molecular masses of around 75 kDa, i.e., p1 (78 kDa), p2 (71 kDa), p3 (74 kDa), and p4 (72 kDa), were inducibly overexpressed in the
Bacterial System for Alginate Uptake and Degradation
81
outer membrane of alginate-grown strain A1 cells (Hashimoto et al. 2005a). The outer-membrane proteins showed a significant sequence identity with TonB-dependent outer-membrane transporters. A mutant with a disruption of the p1, p2, p3, or p4 gene showed a significant growth retardation in the alginate medium, suggesting that each protein was involved in alginate assimilation. The typical outer-membrane transporters so far analyzed incorporate iron-bound siderophores, such as ferrichrome, enterobactin, enterochelin, and citrate (Coulton et al. 1986; Pressler et al. 1988; Lundrigan and Kadner 1986; Koedding et al. 2004) into the periplasm, with a proton motive force generated from an inner-membrane complex (TonB-ExbB-ExbD) (Larsen et al. 1999). Alginate potentially chelates iron (Sreeram et al. 2004), suggesting that p1, p2, p3, and/or p4 may incorporate iron into the periplasm using alginate as a siderophore. Homology modeling indicated that p3 constitutes a tunnel-like b-barrel structure spanning the outer membrane (Fig. 2). After the alginate-metal ion complex had been concentrated by p8 in the pit, the iron-bound alginate was transported across the outer membrane through the action of the TonB-dependent transporters, p1, p2, p3, and/or p4.
3.3
Periplasmic Binding Proteins
In the presence of alginate, strain A1 cells inducibly express two proteins, AlgQ1 (59 kDa) and AlgQ2 (59 kDa), in the periplasm. Both mediate the transfer of alginate from the outer membrane (pit) to the inner membrane (Fig. 2). AlgQ1 and AlgQ2 are mutually similar in primary structure and were found to specifically bind to alginate with a Kd of 2.3 × 10−7 and 1.5 × 10−7 M, respectively (Momma et al. 2005). Neighboring genes code for AlgQ1 and AlgQ2 in the genome of strain A1 (Fig. 2, bottom). X-ray crystallographic studies indicate that the overall structures of AlgQ1 and AlgQ2 are almost identical. Both alginate-binding proteins consist of two N- and C-domains with an a/b structure (Momma et al. 2002, 2005) (Fig. 2). The N- and C-domains are connected by three linker loops, and the cleft formed between the two domains accommodates alginate molecules. Conformational change occurs in alginate-binding proteins through alginate binding and release (Mishima et al. 2003; Momma et al. 2005). In contrast to a ligand-free form whose N- and C-domains are wide open, ligand-bound proteins adopt a closed form (Fig. 3a). Between the open and closed forms, the root-mean-square deviation of the equivalent Ca atoms that were within a distance of 2.0 Å of one another was 0.67 Å (295 Ca atoms). These values were 0.42 Å for the N-domain (224 Ca atoms) and 0.47 Å for the C-domain (264 Ca atoms). The N-domains of the open and closed forms were superimposed, and the rotation and translation required to superimpose the C-domains were determined to be 30° and 0.5 Å, respectively. During the conversion of the open cleft to the closed one, there was a concerted shift of the Glu396 side chain, which moved up into the cleft owing to alginate binding. This ligand-induced movement of Glu396 might be the trigger for the motion that enabled the other
82
W. Hashimoto et al.
a
30˚ C-domain
C-domain C
C
N N
Ca2+
Ca2+ N-domain
N-domain
b 25˚
N C N
C
c
C
Fig. 3 Protein dynamics. a Alginate-binding protein AlgQ2 (left open form, right alginate-bound closed form); b alginate lyase A1-III (left open form, right alginate-bound closed form); c alginate lyase A1-II¢ (left overall structure, right two flexible loops colored blue or red)
domain to participate in ligand binding and ultimately to engulf the bound alginate. The Glu396 movement accompanied the exclusion of a water molecule from the binding site. The major driving force for the hinge closing in alginate-binding proteins is, therefore, the exclusion of a water molecule from the binding site. The shift in the equilibrium on alginate binding might also be enhanced by the
Bacterial System for Alginate Uptake and Degradation
83
interaction of alginate molecules with Glu396, which caused the resulting perturbation of the hinge favoring the closed form. The clefts in AlgQ1 and AlgQ2 were larger than those observed in other substratebinding proteins, such as maltose-binding proteins (Momma et al. 2005, Quiocho et al. 1997). Computational simulation indicated that at least a linear hexasaccharide could be accommodated in the active cleft of alginate-binding proteins, thus demonstrating that this larger space in the cleft enables AlgQ1 and AlgQ2 to bind to macromolecules. Distinct from sugar-binding proteins with aromatic and polar residues located in the cleft (Quiocho et al. 1997), many positively charged residues, as well as aromatic ones, are arranged in the cleft of alginate-binding proteins (Mishima et al. 2003; Momma et al. 2005). Positively charged residues in the active cleft of alginate-binding proteins contribute to the specific binding to acidic polysaccharides, such as alginate. Van der Waals contacts and hydrogen bonds are formed between alginate and binding proteins. The nonreducing terminal residues of alginate molecules are tightly bound to alginate-binding proteins. Thus, AlgQ1 and AlgQ2 specifically recognize and bind to the nonreducing saccharide of alginate, and deliver the polymer to the ATP-binding cassette (ABC) importer in the inner membrane.
3.4
Inner-Membrane ABC Importer
Bacterial ABC importers typically consist of four subunits (two ATP-binding proteins and two transmembrane subunits) (Davidson and Chen 2004). In most Gram-negative bacteria, such as E. coli, ABC importers accompany a periplasmic binding protein mediating the transport of the substrate from the outer membrane to an ABC importer in the inner membrane, and the genes coding for the binding protein and ABC importer are assembled into a cluster (Boos and Shuman 1998). A system consisting of ATP-binding proteins, transmembrane domains, and an associated periplasmic binding protein is thus crucial for substrate import. This is also true in the ABC importer system for alginate in strain A1, i.e., the strain A1 genetic cluster for the alginate import system contains five genes for the ATP-binding proteins (AlgS/AlgS), transmembrane domains (AlgM1 and AlgM2), and two alginate-binding proteins (AlgQ1 and AlgQ2), in that order (Momma et al. 2000) (Fig. 2, bottom). The strain A1 alginate ABC importer has a general architecture composed of two molecules of AlgS (40 kDa) and a heterodimer of AlgM1 (37 kDa) and AlgM2 (33 kDa) (Fig. 2). AlgM1 and AlgM2 are homologous to the permease domain of a bacterial ABC importer. P-loop/Walker and ABC-signature motifs are also conserved in AlgS, which exhibits ATPase activity in the dimeric form. In the alginate ABC importer, AlgM1 and AlgM2 function as permeases for the transport of alginate from the periplasm to the cytoplasm using the energy derived from ATP hydrolysis, catalyzed by the AlgS homodimer. The three-dimensional structure of the alginate ABC importer has been modeled through homology modeling and docking simulation (Fig. 2). AlgM1 and AlgM2 each
84
W. Hashimoto et al.
contain six transmembrane a-helices. A pore for the passage of alginate is formed at the interface between AlgM1 and AlgM2. A crystallographic analysis of AlgS indicates that the protein includes an a/b structure (Mishima et al. 2001). The interface between the transmembrane domains (AlgM1-AlgM2) and ATP-binding proteins (AlgS-AlgS) is stabilized by the interaction of a short a-helix of AlgM1 and AlgM2 extruding to the cytoplasm with an ENI motif and a Q loop of AlgS. Although a dimeric form of AlgS is active as ATPase in vitro, even after separation from the alginate ABC importer, the ATPase activity is substantially increased in the presence of both transmembrane domains (AlgM1 and AlgM2) and periplasmic binding proteins (AlgQ1 and AlgQ2), complexed with alginate (Momma et al. 2005). The activation of AlgS in the dimer form requires the participation of the transmembrane domains (AlgM1 and AlgM2) and alginate-bound proteins (AlgQ1 and AlgQ2), thus suggesting that information on the structural change induced in the binding proteins through alginate binding is transferred to a dimer form of AlgS via mediation by the transmembrane domains. As is seen in the E. coli maltose transport system (Oldham et al. 2007), alginate-bound AlgQ1 or AlgQ2 initiates the transport cycle. Once AlgS has bound to ATP, domains in the alginate-bound AlgQ1/AlgQ2 open wide (release of alginate), and the periplasmic entrance is inducibly produced by the transmembrane domains (AlgM1 and AlgM2). Subsequently, alginate is transported to the cytoplasm by the ABC importer using the energy generated by the ATP hydrolysis. The expression of the alginate ABC importer affects the pit formation on the cell surface (Momma et al. 2000). Mutants with disruption of the ABC importer genes form no apparent pit on the cell surface and fail to assimilate alginate. In conclusion, the clever system used by strain A1 for alginate uptake is composed of a pit on the cell surface, periplasmic binding proteins, and an ABC importer, which function cooperatively.
4
Alginate Degradation
Alginate molecules contain three blocks, polyM, polyG, and polyMG. In strain A1, alginate incorporated into the cytoplasm through the alginate uptake system is degraded to the constituent monosaccharides by endotype and exotype alginate lyases with different substrate (block) specificities.
4.1
Endotype Alginate Lyases
Strain A1 cells produce three endotype alginate lyases in their cytoplasm, A1-I (65 kDa), A1-II (25 kDa), and A1-III (40 kDa) (Yoon et al. 2000) (Fig. 2). These three endotype alginate lyases originate from a single gene (Murata et al. 1993). In the genome of strain A1, this gene (aly) is located in the vicinity of the operon for
Bacterial System for Alginate Uptake and Degradation
85
the alginate uptake system (Fig. 2, bottom). Po (71 kDa) is first synthesized in an inactive form of alginate lyase as a precursor of the gene product. A1-I is converted from Po as an active form after the cleavage of the N-terminal peptide (6 kDa). A1-I is subsequently processed into N-terminal A1-III and C-terminal A1-II in an autocatalytic manner. A1-III and A1-II are specific for polyM and polyG, respectively, indicating that A1-I is a combined form of A1-III and A1-II, and has the characteristics of both enzymes. On the basis of their primary structures, polysaccharide lyases are classified into 18 families (PL-1–PL-18) in the carbohydrate-active enzyme (CAZY) database (B. Henrissat, P. Coutinho, and E. Deleury, http://afmb.cnrs-mrs.fr/~cazy/CAZY/ index.html). Alginate lyases are categorized as PL-5, PL-7, PL-14, PL-15, PL-17, and PL-18. Most bacterial alginate lyases are classified into two families, PL-5 and PL-7, and the alginate lyases in these families are generally specific for polyM and polyG, respectively. Endotypes A1-III and A1-II belong to PL-5 and PL-7, respectively, and thus A1-I becomes a member of a new family PL-5 + 7. Some pseudomonads, such as P. aeruginosa PAO1 (Stover et al. 2000) and Pseudomonas syringae pv. tomato DC3000 (Buell et al. 2003) have both PL-5 and PL-7 alginate lyases. Distinct from strain A1, including the PL-5 + 7 family’s enzyme gene, the PL-5 and PL-7 enzyme genes of pseudomonads are located separately in their bacterial genomes. It is postulated that A1-I is an original alginate lyase, and that the A1-III and A1-II genes derived from the A1-I gene, independently evolving into various genes belonging to the PL-5 and PL-7 families through duplication, modification, and translocation. The genome sequence of strain A1 provides hints to the molecular diversity and evolution of alginate lyases as follows (Miyake et al. 2004). In addition to the genes for A1-I, A1-II, and A1-III, the genome of strain A1 contains a novel alginate lyase (A-II¢, 31 kDa) gene, although no A1-II¢ is expressed in strain A1 cells even in the presence of alginate, and the location of the A1-II¢ gene in the genome of strain A1 is far from the gene cluster for alginate uptake and degradation. A1-II¢ is a member of PL-7, on the basis of the primary structure. Distinct from A1-II, A1-II¢ shows a broad substrate specificity (polyM, polyG, and polyMG). A phylogenetic analysis of the alginate lyases categorized into PL-5 and PL-7 indicated that A1-I has a central position in the phylogenetic tree, being an intermediate between PL-5 and PL-7. This suggested that the A1-II¢ gene was generated from the A1-II gene through separation from the A1-I gene, subsequent duplication, modification, and translocation. Since alginate lyases have potential applications as biochemicals for processing edible seaweed alginate and removing bacterial biofilm alginate, the structure and function relationship of these alginate lyases has been analyzed. An X-ray crystallographic analysis demonstrated that PL-5 A1-III consists of 12 a-helices constituting an a6/a5-barrel structure as a basic frame (Yoon et al. 1999) (Fig. 3b). The enzyme has a tunnel-like cleft covered with a lid loop consisting of residues 57–90. A comparative structural analysis of substrate-free and bound A1-III indicated that the loop is situated about 16.8 Å above the glycosidic linkage to be cleaved in the ligand-free form (open conformation), while this distance is 8.6 Å in the closed formation (Yoon et al. 2001; Mikami et al. 2002). The movement of the lid loop is essential
86
W. Hashimoto et al.
for enzyme catalysis, activation of the catalytic center Tyr246 through the formation of the Tyr68–Tyr246 pair, the binding of substrates, and the release of products. The catalytic reaction of polysaccharide lyases has a three-step mechanism (Linker et al. 1956): (1) the removal or neutralization of the negative charge on the C6 carboxylate anion by residue I; (2) the abstraction of the C-5 proton at subsite + 1 by a general base (residue II); and (3) proton donation to the O4 atom at subsite + 1 by a general acid (residue III). The alginate tetrasaccharide-bound structure of A1-III indicates that His192 functions as residue I, and Tyr246 as both residues II and III. The catalysis by A1-III with a single Tyr as an active center can probably be regarded as a novel b-elimination reaction. X-ray crystallographic studies revealed that PL-7 A1-II¢, which had a broad substrate specificity, consists of two b-sheets constituting a b-sandwich structure. A1-II¢ has an active site in the cleft covered with two flexible loops (residues 133–145 and 193–203) (Yamasaki et al. 2005). The flexibility of these loops is essential for substrate binding, i.e., the accommodation of substrates in the active site (Ogura et al. 2008) (Fig. 3c). Both G and M in alginate molecules are bound to the active site as follows: the enzyme interacts appropriately with the substrate’s hydroxyl groups, although the substrate’s carboxyl groups are recognized by specific residues. These structural characteristics provide A1-II¢ with its broad substrate specificity. In the A1-II¢ reaction, Gln189 (residue I) neutralizes the negative charge of the carboxyl group, His191 (residue II) abstracts the C-5 proton, and Tyr284 (residue III) donates the proton to the glycoside bond to be cleaved.
4.2
Exotype Alginate Lyases
Alginate is depolymerized to unsaturated disaccharides, trisaccharides, and tetrasaccharides with different M to G ratios as final products through the reaction of the endotype alginate lyases A1-I, A1-II, and A1-III (Yoon et al. 2000). Alginate oligosaccharides are further exolytically degraded to the constituent monosaccharides by alginate lyase A1-IV (86 kDa) (Hashimoto et al. 2000). The resultant monosaccharides are nonenzymatically converted to an a-keto acid (4-deoxy-l-erythro-5-hexoseulose uronic acid). The genes coding for A1-I, A1-II, A1-III, and A1-IV are included in the genetic cluster for the system for alginate uptake and degradation (Murata et al. 1993; Momma et al. 2000; Hashimoto et al. 2000) (Fig. 2, bottom), and this genetic cluster was found to be inducibly expressed in the presence of alginate through transcriptomics by DNA microarray. Therefore, the alginate uptake system is linked with the degradation system, and the expression is coordinately regulated. A1-IV exolytically acts on alginate polysaccharides and oligosaccharides in saturated form, as well as on unsaturated oligosaccharides, indicating that A1-IV can degrade saturated and unsaturated alginate polysaccharides and oligosaccharides with various M and G compositions (Miyake et al. 2003). A1-IV is, therefore, involved in the complete degradation of alginate in strain A1. Exotype A1-IV, with its broad substrate specificity, is peculiar among alginate lyases in that the enzymes so far
Bacterial System for Alginate Uptake and Degradation
87
analyzed prefer either polyM or polyG, and act on alginate endolytically, although there have been a few reports on alginate lyases that depolymerize both polyM and polyG, and show exolytic activity (Wong et al. 2000). To the best of our knowledge, no alginate lyase has previously been reported that exolytically acts on unsaturated and saturated alginate molecules of various lengths and M to G ratios, except for A1-IV. This is why A1-IV shows no significant identity with any other functionknown proteins. Accompanying the bacterial genome sequence analysis, novel genes encoding hypothetical proteins homologous with A1-IV were found in the genomes of some bacteria, including strain A1. One such protein, A1-IV¢ (90 kDa) of strain A1, catalyzes the cleavage of glycosidic bonds in alginate molecules through the b-elimination reaction, and releases unsaturated disaccharides and trisaccharides as the main products, indicating that the enzyme is an endotype alginate lyase (Hashimoto et al. 2005b). On the other hand, Agrobacterium tumefaciens Atu3025 (88 kDa), which is similar to A1-IV, was identified as an exotype alginate lyase (Ochiai et al. 2006). On the basis of their primary structures, A1-IV and other A1-IV homologues contribute to the establishment of a novel polysaccharide lyase family, PL-15, suggesting the evolutionary route of alginate lyases in PL-15 (Hashimoto et al. 2005b). As was seen in the case of A1-II¢, A1-IV¢ was not expressed in strain A1 cells grown on alginate. It is thought that the A1-IV¢ gene was generated from the A1-IV gene and that A1-IV was the original protein of PL-15. In conclusion, strain A1 produces diverse alginate lyases classified into the families PL-5, PL-7, PL-5 + 7, and PL-15, which contain almost all bacterial alginate lyases, suggesting the possibility that strain A1 can be regarded as the original producer of alginate lyases.
5
Concluding Remarks and Future Perspectives
A large number of genome sequencing projects are now being carried out, and the complete gene sequences of over 700 organisms, ranging from viruses to human, have been determined. Several “omics” technologies, such as transcriptomics, proteomics, and glycomics, have facilitated the interpretation of the enormous amount of information obtained in genome sequencing. Comprehensive analyses of genes, proteins, and saccharides using these technologies readily and simultaneously contribute to our understanding of interesting strain A1 cellular events, especially from the pattern of global gene expression.
5.1
Superchannel as Alginate Uptake System
When microbes utilize macromolecules for their growth, they usually produce extracellular macromolecule-degrading enzymes, and incorporate the degraded low molecular weight products through specific transporters, channels, and/or porins. On the other
88
W. Hashimoto et al.
hand, strain A1 employs an alternative elegant means for the degradation of macromolecules (Fig. 2) (Murata et al. 2008). Once cell surface receptors (p5 and p6) recognize the external alginate, the resultant signals are transmitted to the DNA. On the basis of signal transduction, genes responsible for alginate uptake and degradation are expressed. Alginate is concentrated in the cell surface pit where alginate-binding proteins (p7 and p8) are inducibly expressed in the presence of alginate, and is then incorporated by TonB-dependent transporters (p1–p4). Periplasmic alginate-binding proteins (AlgQ1 and AlgQ2) deliver the polysaccharide from the outer membrane to the inner membrane. The inner-membrane ABC importer (AlgM1–AlgM2/AlgS–AlgS) directly incorporates alginate into the cytoplasm with the energy generated by AlgS. This is a novel macromolecule uptake system independent of extracellular degrading enzymes, which we call a “superchannel.” This assimilation method (swallow type) for macromolecules is energetically and substantially more efficient and economical than the extracellular-degrading-enzyme-dependent assimilation method (chewing type), since it minimizes the degraded product loss from diffusion that is observed in the chewing-type method. The significance of the novel pit-dependent ABC importer system for macromolecule uptake is that cell surface molecules are in a fluidized state and are rearranged to create a molecular apparatus. Similar to the strain A1 superchannel, other pit or pore-dependent transport systems are found in bacterial and human cells (Sugawara et al. 1999; Andrews et al. 1990; Noronha et al. 2000; Shimizu 1999). This pit formation may be considered to be the original model for the endocytosis and/or phagocytosis of eukaryotic cells, and this notion can be applied to the uptake of other macromolecules across the cell membrane, such as DNA and proteins.
5.2
Molecular Transplantation of Superchannel
An understanding of the direct uptake system, i.e., superchannel, present in strain A1 is not only beneficial in clarifying the macromolecule transport mechanism in microbial cells, but also has a potential application for improving in the permeability of various compounds across cell membranes. The widespread use of chemicals, such as solvents, herbicides, insecticides, and many industrial chemicals, has led to the contamination of soils, waste streams, and various other sites, and has directly or indirectly had hazardous effects on many organisms through food chains. Among these chemicals, dibenzofuran, a dioxin analogue, is a highly toxic compound generated in combustion processes and certain chemical syntheses of haloaromatics. To eliminate it, as well as other hazardous compounds, several physicochemical processes have been considered, such as hightemperature incineration, but these methods are not always economically feasible for the treatment of pollutants, which are present at low concentrations and in soluble/insoluble forms. More efficient and convenient biological methods, using microorganisms, are now sought to assist in their removal from contaminated sites. At present, most bioremediation studies are restricted to the isolation of microbes capable of degrading environmental pollutants and analyses of their genes/enzymes
Bacterial System for Alginate Uptake and Degradation
89
involved in the degradation pathway. Among the pollutant-degrading microbes so far isolated and analyzed, sphingomonads are the most promising microbes for potential bioremediation applications. Sphingomonads are widely distributed in nature and are frequently found in various soils and aqueous environments. Sphingomonas wittichii RW1 (strain RW1) shows significant growth using dibenzofuran as a source of carbon and energy (Wittich et al. 1992). To enhance the permeability of the molecules, the molecular transplantation of the strain A1 superchannel to strain RW1 was conducted through the introduction of the genetic cluster for the alginate uptake system (Aso et al. 2006). The strain RW1 transformant with the genetic cluster formed a pit on the cell surface, while no pit was observed in the wild type of strain RW1 (Fig. 4). The wild type of strain RW1 showed a lag time during the first few days, and a longer cultivation time (4 days) was required for the complete removal of dibenzofuran. In comparison with the wild type of strain RW1, the transformant degraded the dibenzofuran rapidly, within 2 days. After being given the strain A1 genetic cluster, the poly(propylene glycol) degrading Sphingomonas subarctica and polydextrose-degrading Sphingomonas sanguis also formed the pit constitutively on their cell surface and showed significant macromoleculedegrading abilities (Aso et al. 2006). On the basis of the results described above, the molecular transplantation of the superchannel has been established as a novel practical biotechnology for the molecular breeding of microbial degraders suitable for the bioremediation of various environmental pollutants with low to high molecular weights.
Dioxin
Sphingomonas sp. A1 Genes for superchannel Plasmid
Superbacteria
S. wittichii RW1 (wild-type)
S. wittichii RW1 (transformant)
Fig. 4 Molecular transplantation of the superchannel
90
5.3
W. Hashimoto et al.
Application of the Alginate Degradation System
Alginate degradation products are expected to have applications in various food, chemical, and medical formulations because of the safety record of these polymers. A wide range of industries have long been interested in alginate-modifying enzymes. The use of alginate lyases will enable the preparation of physiologically and technologically novel and available materials in the agricultural, food, and pharmaceutical fields. The results presented in this chapter, therefore, provide a possible and promising way, when taken together, of obtaining a more complete picture of the utilization of alginate, as well as of the renewable oligosaccharides that are obtainable in large quantities at low cost. In the genome of strain A1, several genes coding for alginate lyases with different substrate specificities and modes of action were found. These enzymes are promising for removing and modifying alginate. Pathogenic bacteria, such as P. aeruginosa, are known to produce exopolysaccharides as a capsulelike biofilm that facilitates interaction with and complete adherence to target cells (Costerton et al. 1999). This biofilm often forms a firm barrier and inhibits the penetration of antimicrobial agents, making it difficult to treat biofilm-dependent bacterial infections. Degradation of the bacterial exopolysaccharides might lead to the establishment of a novel therapy for biofilm-dependent bacterial infectious diseases. The application of A1-III, with its high activity for Pseudomonas alginate, might be feasible for the treatment of such diseases (Sakakibara et al. 2002). The physiological function of oligosaccharides is of importance. Alginate oligosaccharides depolymerized with A1-I enhance the proliferation and/or differentiation of bifidobacteria (Akiyama et al. 1992) and some plant cells (Murata et al. 1993), making it possible to use alginate oligosaccharides as growth factors for bifidobacteria in the food industry or as biochemical fertilizers in agriculture. The alginate oligosaccharides obtained with A1-I also promote the differentiation and proliferation of human epithelium cells (Hashimoto et al. 2002), indicating that these oligosaccharides might be applicable as epidermal growth factors in place of bovine pituitary extract, the utilization of which is thought to be hazardous owing to the occurrence of bovine spongiform encephalopathy (Kawada et al. 1997). Biofuel is expected to become a new energy source in place of fossil fuels, such as petroleum and coal (Gray et al. 2006). In particular, technologies for the production of bioethanol are sought and the industrial production of ethanol from sugarcane and corn starch has been developed (Gray et al. 2006). The use of foodstuffs to produce ethanol, however, causes serious social problems, such as insufficient supply and a sudden rise in prices for the crops utilized. Lignocellulose is thought to be the next target for bioethanol production, but some technical problems regarding how to remove the lignin and hydrolyze the cellulose still need to be overcome. Recently, marine biomass has been focused on as a possible material for bioethanol production. Alginate, a major part of this marine biomass, makes up 30–60% of dried seaweed. The strain A1 exotype alginate lyase A1-IV, with its broad substrate specificity, degrades alginate to the constituent monosaccharides, indicating that it is feasible to use this enzyme for the saccharification of alginate. The a-keto acid
Bacterial System for Alginate Uptake and Degradation
91
derived from the resultant monosaccharides is further converted to pyruvate and glyceraldehyde 3-phosphate through successive reactions of cytoplasmic reductase A1-R (28 kDa), kinase A1-K (35 kDa), and aldolase A1-A (24 kDa) (Ochiai et al. 2008). This metabolic pathway has also been observed in P. aeruginosa (Preiss and Ashwell 1962). The tricarbon chemicals pyruvate and glyceraldehyde 3-phosphate can be converted into ethanol. In fact, strain A1 cells grown on alginate under a limited oxygen level synthesize ethanol, suggesting that strain A1 may become a potent ethanologen through assimilating alginate, although the production level of ethanol should be enhanced through metabolic engineering. Acknowledgements This work was supported by Grants-in-Aid (to K.M., B.M., and W.H.) and by the Targeted Proteins Research Program (to W.H.) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan. Part of this work was also supported by the Program of Basic Research Activities for Innovative Biosciences (PROBRAIN) of Japan (to K.M.).
References Aizawa S-I (1996) Flageller assembly in Salmonella typhimurium. Mol Microbiol 19:1–5 Akiyama H, Endo T, Nakakita R, Murata K, Yonemoto Y, Okayama K (1992) Effect of depolymerized alginates on the growth of bifidobacteria. Biosci Biotechnol Biochem 56:355–356 Andrews NW, Abrams CK, Slatin SL, Griffiths G (1990) A T. cruzi-secreted protein immunologically related to the complement component C9: evidence for membrane pore-forming activity at low pH. Cell 61:1277–1287 Aso Y, Miyamoto Y, Harada KM, Momma K, Kawai S, Hashimoto W, Mikami B, Murata K (2006) Engineered membrane superchannel improves bioremediation potential of dioxin-degrading bacteria. Nat Biotechnol 24:188–189 Boos W, Shuman H (1998) Maltose/maltodextrin system of Escherichia coli: transport, metabolism, and regulation. Microbiol Mol Biol Rev 62:204–229 Buell CR, Joardar V, Lindeberg M, Selengut J, Paulsen IT, Gwinn ML, Dodson RJ, Deboy RT, Durkin AS, Kolonay JF, Madupu R, Daugherty S, Brinkac L, Beanan MJ, Haft DH, Nelson WC, Davidsen T, Zafar N, Zhou L, Liu J, Yuan Q, Khouri H, Fedorova N, Tran B, Russell D, Berry K, Utterback T, Van Aken SE, Feldblyum TV, D’Ascenzo M, Deng WL, Ramos AR, Alfano JR, Cartinhour S, Chatterjee AK, Delaney TP, Lazarowitz SG, Martin GB, Schneider DJ, Tang X, Bender CL, White O, Fraser CM, Collmer A (2003) The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc Natl Acad Sci U S A 100:10181–10186 Costerton JW, Stewart PS, Greenberg EP (1999) Bacterial biofilms: a common cause of persistent infections. Science 284:1318–1322 Coulton JW, Mason P, Cameron DR, Carmel G, Jean R, Rode HN (1986) Protein fusions of b-galactosidase to the ferrichrome-iron receptor of Escherichia coli K-12. J Bacteriol 165:181–192 Davidson AL, Chen J (2004) ATP-binding cassette transporters in bacteria. Annu Rev Biochem 73:241–268 Gacesa P (1988) Alginates. Carbohydr Polym 8:161–182 Gray KA, Zhao L, Emptage M (2006) Bioethanol. Curr Opin Chem Biol 10:141–146 Hanley SZ, Pappin DJ, Rahman D, White AJ, Elborough KM, Slabas AR (1999) Re-evaluation of the primary structure of Ralstonia eutropha phasin and implications for polyhydroxyalkanoic acid granule binding. FEBS Lett 447:99–105 Hashimoto W, Miyake O, Momma K, Kawai S, Murata K (2000) Molecular identification of oligoalginate lyase of Sphingomonas sp. strain A1 as one of the enzymes required for complete depolymerization of alginate. J Bacteriol 182:4572–4577
92
W. Hashimoto et al.
Hashimoto W, Mishima Y, Miyake O, Nankai H, Momma K, Murata K (2002) Biodegradation of alginate, xanthan, and gellan. In: Steinbüchel, A Biopolymers, polysaccharides I, vol 7. Wiley, Weinheim, pp 175–199 Hashimoto W, He J, Wada Y, Nankai H, Mikami B, Murata K (2005a) Proteomics-based identification of outer-membrane proteins responsible for import of macromolecules in Sphingomonas sp. A1: alginate-binding flagellin on the cell surface. Biochemistry 44:13783–13794 Hashimoto W, Miyake O, Ochiai A, Murata K (2005b) Molecular identification of Sphingomonas sp. A1 alginate lyase (A1-IV¢) as a member of novel polysaccharide lyase family 15 and implications in alginate lyase evolution. J Biosci Bioeng 99:48–54 Hashimoto W, Momma K, Maruyama Y, Yamasaki M, Mikami B, Murata K (2005c) Structure and function of bacterial super-biosystem responsible for import and depolymerization of macromolecules. Biosci Biotechnol Biochem 69:673–692 Harada KM, Aso Y, Hashimoto W, Mikami B, Murata K (2006) Sequence and analysis of the 46.6-kb plasmid pA1 from Sphingomonas sp. A1 that corresponds to the typical IncP-1b plasmid backbone without any accessory gene. Plasmid 56:11–23 Hayashi F, Smith KD, Ozinsky A, Hawn TR, Yi EC, Goodlett DR, Eng JK, Akira S, Underhill DM, Aderem A (2001) The innate immune response to bacterial flagellin is mediated by Toll-like receptor 5. Nature 410:1099–1103 He J, Ochiai A, Fukuda Y, Hashimoto W, Murata K (2008) A putative lipoprotein of Sphingomonas sp. A1 binds alginate rather than a lipid moiety. FEMS Microbiol Lett 288:221–226 He J, Nankai H, Hashimoto W, Murata K (2004) Molecular identification and characterization of an alginate-binding protein on the cell surface of Sphingomonas sp. A1. Biochem Biophys Res Commun 322:712–717 Hisano T, Kimura N, Hashimoto W, Murata K (1996) Pit structure on bacterial cell surface. Biochem Biophys Res Commun 220:979–982 Iwamoto M, Kurachi M, Nakashima T, Kim D, Yamaguchi K, Oda T, Iwamoto Y, Muramatsu T (2005) Structure-activity relationship of alginate oligosaccharides in the induction of cytokine production from RAW264.7 cells. FEBS Lett 579:4423–4429 Iwamoto Y, Xu X, Tamura T, Oda T, Muramatsu T (2003) Enzymatically depolymerized alginate oligomers that cause cytotoxic cytokine production in human mononuclear cells Biosci Biotechnol Biochem 67:258–263 Jensen A (1993) Present and future needs for algae and algal polysaccharides. Hydrobiologica 260–261:15 Kawada A, Hiura N, Shiraiwa M, Tajima S, Hiruma M, Hara K, Ishibashi A, Takahara H (1997) Stimulation of human keratinocyte growth by alginate oligosaccharides, a possible co-factor for epidermal growth factor in cell culture. FEBS Lett 408:43–46 Kawada A, Hiura N, Tajima S, Takahara H (1999) Alginate oligosaccharides stimulate VEGFmediated growth and migration of human endothelial cells. Arch Dermatol Res 291:542–547 Kawasaki S, Moriguchi R, Sekiya K, Nakai T, Ono E, Kume K, Kawahara K (1994) The cell envelope structure of the lipopolysaccharide-lacking gram-negative bacterium Sphingomonas paucimobilis. J Bacteriol 176:284–290 Kinjo Y, Wu D, Kim G, Xing GW, Poles MA, Ho DD, Tsuji M, Kawahara K, Wong CH, Kronenberg M (2005) Recognition of bacterial glycosphingolipids by natural killer T cells. Nature 434:520–525 Koedding J, Howard P, Kaufmann L, Polzer P, Lustig A, Welte W (2004) Dimerization of TonB is not essential for its binding to the outer membrane siderophore receptor FhuA of Escherichia coli. J Biol Chem 279:9978–9986 Kurachi M, Nakashima T, Miyajima C, Iwamoto Y, Muramatsu T, Yamaguchi K, Oda T (2005) Comparison of the activities of various alginates to induce TNF-a secretion in RAW264.7 cells. J Infect Chemother 11:199–203 Larsen RA, Thomas MG, Postle K (1999) Proton motive force, ExbB and ligand-bound FepA drive conformational changes in TonB. Mol Microbiol 31:1809–18024 Leiman PG, Chipman PR, Kostyuchenko VA, Mesyanzhinov VV, Rossmann MG (2004) Threedimensional rearrangement of proteins in the tail of bacteriophage T4 on infection of its host. Cell 118:419–429
Bacterial System for Alginate Uptake and Degradation
93
Linker A, Meyer K, Hoffman P (1956) The production of unsaturated uronides by bacterial hyaluronidases. J Biol Chem 219:13–25 Llanes F, Ryan DH, Marchessault RH (2000) Magnetic nanostructured composites using alginates of different M/G ratios as polymeric matrix. Int J Biol Macromol 27:35–40 Lundrigan MD, Kadner RJ (1986) Nucleotide sequence of the gene for the ferrienterochelin receptor FepA in Escherichia coli. Homology among outer membrane receptors that interact with TonB. J Biol Chem 261:10797–10801 Maruyama Y, Momma M, Mikami B, Hashimoto W, Murata K (2008) Crystal structure of a novel bacterial cell-surface flagellin binding to a polysaccharide. Biochemistry 47:1393–1402 May TB, Chakrabarty AM (1994) Pseudomonas aeruginosa: genes and enzymes of alginate synthesis. Trends Microbiol 2:151–157 Mikami B, Suzuki S, Yoon H-J, Miyake O, Hashimoto W, Murata K (2002) X-ray structural analysis of alginate lyase A1-III mutants/substrate complexes: activation of a catalytic tyrosine residue by a flexible lid loop. Acta Crystallogr A 58:C271 Mishima Y, Momma K, Hashimoto W, Mikami B, Murata K (2001) Crystallization and preliminary X-ray analysis of AlgS, a bacterial ATP-binding-cassette (ABC) protein specific to macromolecule import. Acta Crystallogr D Biol Crystallogr 57:884–885 Mishima Y, Momma K, Hashimoto W, Mikami B, Murata K (2003) Crystal structure of AlgQ2, a macromolecule (alginate)-binding protein of Sphingomonas sp. A1, complexed with an alginate tetrasaccharide at 1.6-Å resolution. J Biol Chem 278:6552–6559 Miyake O, Hashimoto W, Murata K (2003) An exotype alginate lyase in Sphingomonas sp. A1: overexpression in Escherichia coli, purification, and characterization of alginate lyase IV (A1-IV). Protein Expr Purif 29:33–41 Miyake O, Ochiai A, Hashimoto W, Murata K (2004) Origin and diversity of alginate lyases of families PL-5 and -7 in Sphingomonas sp. strain A1. J Bacteriol 186:2891–2896 Momma K, Mikami B, Mishima Y, Hashimoto W, Murata K (2002) Crystal structure of AlgQ2, a macromolecule (alginate)-binding protein of Sphingomonas sp. A1 at 2.0Å resolution. J Mol Biol 316:1061–1069 Momma K, Mishima Y, Hashimoto W, Mikami B, Murata K (2005) Direct evidence for Sphingomonas sp. A1: periplasmic proteins as macromolecule-binding proteins associated with ABC transporter: molecular insights into alginate transport in the periplasm. Biochemistry 44:5053–5064 Momma K, Okamoto M, Mishima Y, Mori S, Hashimoto W, Murata K (2000) A novel bacterial ATP-binding cassette (ABC) transporter system that allows uptake of macromolecules. J Bacteriol 182:3998–4004 Moreno S, Nájera R, Guzmán J, Soberón-Chávez G, Espín G (1998) Role of alternative sigma factor algU in encystment of Azotobacter vinelandii. J Bacteriol 180:2766–2769 Murata K, Inose T, Hisano T, Abe S, Yonemoto Y, Yamashita T, Takagi M, Sakaguchi K, Kimura A, Imanaka T (1993) Bacterial alginate lyase: enzymology, genetics and application. J Ferment Bioeng 76:427–437 Murata K, Kawai S, Mikami B, Hashimoto W (2008) Superchannel of bacteria: biological significance and new horizons. Biosci Biotechnol Biochem 72:265–277 Noronha FS, Cruz JS, Beirao PS, Horta MF (2000) Macrophage damage by Leishmania amazonensis cytolysin: evidence of pore formation on cell membrane. Infect Immun 68:4578–4584 Ochiai A, Hashimoto W, Murata K (2006) A biosystem for alginate metabolism in Agrobacterium tumefaciens strain C58: molecular identification of Atu3025 as an exotype family PL-15 alginate lyase. Res Microbiol 157:642–649 Ochiai A, Takase R, Hashimoto W, Murata K (2008) Molecular identification of reductase involved in detoxifying alpha-keto acid derived from alginate monosaccharide. Abstract for the annual meeting of the Society for Biotechnology, Japan, page 182 Ogura K, Yamasaki M, Mikami B, Hashimoto W, Murata K (2008) Substrate recognition by family 7 alginate lyase from Sphingomonas sp. A1. J Mol Biol 380:373–385 Oldham ML, Khare D, Quiocho FA, Davidson AL, Chen J (2007) Crystal structure of a catalytic intermediate of the maltose transporter. Nature 450:515–521
94
W. Hashimoto et al.
Preiss J, Ashwell G (1962) Alginic acid metabolism in bacteria. I. Enzymatic formation of unsaturated oligosaccharides and 4-deoxy-L-erythro-5-hexoseulose uronic acid. J Biol Chem 237:309–316 Pressler U, Staudenmaier H, Zimmermann L, Braun V (1988) Genetics of the iron dicitrate transport system of Escherichia coli. J Bacteriol 170:2716–2724 Quiocho FA, Spurlino JC, Rodseth LE (1997) Extensive features of tight oligosaccharide binding revealed in high-resolution structures of the maltodextrin transport/chemosensory receptor. Structure 5:997–1015 Remminghorst U, Rehm BH (2006) Bacterial alginates: from biosynthesis to applications. Biotechnol Lett 28:1701–1712 Sakakibara H, Tamura T, Suzuki T, Hisano T, Abe S, Murata K (2002) Preparation and properties of alginate lyase modified with poly(ethylene glycol). J Pharm Sci 91:1191–1199 Schweizer HP, Boring III JR (1973) Antiphagocytic effect of slime from a mucoid strain of Pseudomonas aeruginosa. Infect Immun 3:762–767 Shimizu M (1999) Modulation of intestinal functions by food substances. Nahrung 43:154–158 Sreeram KJ, Shrivastava HY, Nair BU (2004) Studies on the nature of interaction of iron(III) with alginates. Biochim Biophys Acta 1670:121–125 Stover CK, Pham XQ, Erwin AL, Mizoguchi SD, Warrener P, Hickey MJ, Brinkman FSL, Hufnagle WO, Kowalik DJ, Lagrou M, Garber RL, Goltry L, Tolentino E, Westbrock-Wadman S, Yuan Y, Brody LL, Coulter SN, Folger KR, Kas A, Larbig K, Lim R, Smith K, Spencer D, Wong GK-S, Wu Z, Paulsen IT, Reizer J, Saier MH, Hancock REW, Lory S, Olson MV (2000) Complete genome sequence of Pseudomonas aeruginosa PAO1, an opportunistic pathogen. Nature 406:959–964 Sugawara N, Tomita T, Sato T, Kamio Y (1999) Assembly of Staphylococcus aureus leukocidin into a pore-forming ring-shaped oligomer on human polymorphonuclear leukocytes and rabbit erythrocytes. Biosci Biotechnol Biochem 63:884–891 Takeuchi M, Hamana K, Hiraishi A (2001) Proposal of the genus Sphingomonas sensu stricto and three new genera, Sphingobium, Novosphingobium and Sphingopyxis, on the basis of phylogenetic and chemotaxonomic analyses. Int J Syst Evol Microbiol 51:1405–1417 Turley EV, Noble PW, Bourguignon LYW (2002) Signaling properties of hyaluronan receptors. J Biol Chem 277:4589–4592 Walker JC (1987) Was the Archaean biosphere upside down? Nature 329:710–712 White DC, Sutton SD, Ringelberg DB (1996) The genus Sphingomonas: physiology and ecology. Curr Opin Biotechnol 7:301–306 Wittich RM, Wilkes H, Sinnwell V, Francke W, Fortnagel P (1992) Metabolism of dibenzo-p-dioxin by Sphingomonas sp. strain RW1. Appl Environ Microbiol 58:1005–1010 Wong TY, Preston LA, Schiller NL (2000) Alginate lyase: review of major sources and enzyme characteristics, structure-function analysis, biological roles, and applications. Annu Rev Microbiol 54:289–340 Yamasaki M, Ogura K, Hashimoto W, Mikami B, Murata K (2005) A structural basis for depolymerization of alginate by polysaccharide lyase family-7. J Mol Biol 352:11–21 Yonekura K, Maki-Yonekura S, Namba K (2003) Complete atomic model of the bacterial flagellar filament by electron cryomicroscopy. Nature 424:643–650 Yoon H-J, Mikami B, Hashimoto W, Murata K (1999) Crystal structure of alginate lyase A1-III from Sphingomonas species A1 at 1.78 Å resolution. J Mol Biol 290:505–514 Yoon H-J, Hashimoto W, Miyake O, Okamoto M, Mikami B, Murata K (2000) Overexpression in Escherichia coli, purification, and characterization of Sphingomonas sp. A1 alginate lyases. Protein Expr Purif 19:84–90 Yoon H-J, Hashimoto W, Miyake O, Murata K, Mikami B (2001) Crystal structure of alginate lyase A1-III complexed with trisaccharide product at 2.0 Å resolution. J Mol Biol 307:9–16
Enzymatic Alginate Modification Helga Ertesvåg, Svein Valla, and Gudmund Skjåk-Bræk
Contents 1 2
Introduction........................................................................................................................ Mannuronan C-5-Epimerases ............................................................................................ 2.1 Early Studies on the Secreted Mannuronan C-5-Epimerase from A. vinelandii ..................................................................................................... 2.2 Cloning and Characterization of the Mannuronan C-5-Epimerases from A. vinelandii ..................................................................................................... 2.3 Cloning and Characterization of the Bacterial Periplasmic Mannuronan C-5-Epimerases ................................................................................... 2.4 Mannuronan C-5-Epimerases from Algae ................................................................ 2.5 Structure of the Mannuronan C-5-Epimerases ......................................................... 2.6 Mode of Action of the Secreted Mannuronan C-5-Epimerases ................................ 2.7 Enzyme Specificity; Epimerization of Nonnatural Uronans and Poly(guluronic acid) ................................................................................................. 2.8 In Vitro Epimerization with AlgE Epimerases ......................................................... 3 Alginate Lyases .................................................................................................................. 3.1 Alginate Lyases as Tools To Determine Alginate Structure ..................................... 3.2 An Alginate Lyase Is Needed To Remove Aberrantly Localized Alginate .............. 4 Alginate Acetylation .......................................................................................................... 5 Use of Alginate-Modifying Enzymes for Biotechnological Purposes............................... References ................................................................................................................................
96 98 98 100 102 103 103 104 106 107 108 108 108 109 109 110
Abstract Alginate is a linear 1-4-linked copolymer of b-d-mannuronic acid and its C-5-epimer a-l-guluronic acid. The polymer is produced by some algae and bacteria, and is used for numerous purposes in industry. Alginate is initially synthesized as mannuronan, which is then modified at the polymer level by mannuronan C-5-epimerases, alginate lyases, and O-acetylases. This generates a variety of heteropolymers where properties such as viscosity, chain stiffness, gel formation, water-binding potential, and immunogenicity are dependent on the action of the H. Ertesvåg( ), S. Valla and G. Skjåk-Bræk Department of Biotechnology, Norwegian University of Science and Technology, Sem Sælands vei 6/8, 7491 Trondheim, Norway e-mail: [email protected], e-mail: [email protected], e-mail: [email protected] B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_4, © Springer-Verlag Berlin Heidelberg 2009
95
96
H. Ertesvåg et al.
modifying enzymes. Both alginate lyases and C-5-epimerases can be used in vitro to tailor alginates for specific purposes. The lyases may also be used as tools to better define the sugar monomer sequences of an alginate sample.
1
Introduction
Alginates are structurally heterogenous polysaccharides produced by bacteria belonging to the genera Pseudomonas and Azotobacter, and by brown seaweeds and some calcareous red algae (Stanford 1883; Gorin and Spencer 1966; Linker and Jones 1966; Govan et al. 1981; Okazaki et al. 1982). In bacteria these polymers probably serve many different functions, as illustrated, for example, by P. aeruginosa, where they provide the producing organism with a selective advantage during serious infections of patients suffering from the disease cystic fibrosis (Ramsey and Wozniak 2005). In A. vinelandii they play a critical role in the ability of the organism to enter a desiccation-resistant state designated “cyst” (Sadoff 1975). Brown seaweed alginates serve a role in maintaining the integrity of the cell wall, analogous to cellulose and pectins in green land plants, and provide the stiffness or elasticity needed depending on the environment (Haug et al. 1974). The function in red algae is not known although it has been suggested that alginate is involved in the calcification process (Bilan and Usov 2001). All alginates isolated from natural sources are linear 1-4-linked copolymers of the two sugar monomers b-d-mannuronic acid (M) and its C-5-epimer a-l-guluronic acid (G) (Fig. 1). However, alginate structures can otherwise be quite heterogeneous, even when they are isolated from different parts of the same organism (e.g., seaweed blades and stipes) or from bacteria exposed to different environmental conditions. The structural differences relate to the following: (1) the ratio between M and G residues may vary over a wide range (Ertesvåg et al. 1996); (2) the distribution pattern of the monomers along the polymer chains may vary drastically, even if the fractional composition is similar (Skjåk-Bræk et al. 1986a); (3) bacterial alginates are O-acetylated to varying degrees in the 2 and/or 3 positions, while this is not seen in seaweed alginates (Skjåk-Bræk et al. 1986a); (4) alginate chain lengths vary and are heterogenous even when alginate is extracted from a single source. All these four structural differences have very important consequences for the physicochemical properties of the polymer, and the producing organisms have taken advantage of the structural diversity to serve their functional needs. The diversity also forms the basis for a widespread and varied use of the polymer in industry. Generally, the presence of G residues in alginates opens the potential for formation of polymer gels in the presence of divalent cations such as Ca2+, but this form of gel formation can only take place if the G residues are found as consecutive stretches (designated “G-blocks”) (Fig. 1). The lengths and numbers of these blocks affect several important physical properties of the gels, such as stiffness, swelling, and porosity. The G residues may also be present as alternating sequences, (MG)n, designated “MG-blocks” (Fig. 1). Finally, alginates with a very low content of G
Enzymatic Alginate Modification OH O
HOOC O HO
HO O O HOOC OH
97 HOOC O HO
HO O O HOOC OH
M-block
OH O
HOOC O HO
HO O O HOOC OH
OH O
MG-block forming C-5 epimerase i.e. AlgE4
HOOC HO O O HOOC OH
O
OH
HOOC
HO O O HOOC OH OH
O
O
HOOC HO O O HOOC OH
OH O
O
OH G-block-forming C-5 epimerase i.e. AlgE6 OH
MG-block
OH HOOC
HOOC O
OH O
OH O OH
O O
O
OH O
OH HOOC
OH
O O
OH HOOC O
OH OH O
OH HOOC
OH
O O
OH HOOC
G-block
Fig. 1 Alginate structure. Hexamers of the three different structural elements are shown. The mannuronan C-5-epimerases will convert consecutive stretches of b-d-mannuronic acid (M-blocks) to blocks containing single a-l-guluronic acid residues or consecutive stretches of a-l-guluronic acid residues (G-blocks)
residues are referred to as being rich in M-blocks (Fig. 1). Such alginates cannot form cationic gels, but can form acidic gels (Draget et al. 1994). For further details on the relation between monomer composition and physicochemical properties, see Smidsrød and Draget (1996). An additional interesting observation is that some types of alginates exhibit immune-stimulating properties. These activities which include stimulation of cytokine synthesis through the Toll-like receptor system, is correlated with a high content of M interspaced with single G residues, although the sequential structure of the epitope remains to be identified (Flo et al. 2000). The biotechnological uses of alginates cover a wide range of applications owing to the consequences of the structural diversity (Smidsrød and Draget 1996). They are, for example, used to affect food texture owing to the ability to form viscous solutions (related to polymer chain lengths), for immobilization of cells owing to the ability to form cationic gels (Draget et al. 2000), and as a feed additive in fish farming owing to their immunogenic properties (http://www.biomar.no). Since all commercial alginates are currently harvested from brown seaweeds, different structural types of alginates must be selected from different types of alga or from specific algal tissues, i.e., alginate enriched in G for making mechanically strong gels is found in the stipes and holdfast of Laminaria hyperborea, a plant growing in very exposed coastal areas. Alginates with a high M content (for softer and more elastic gels) can be extracted from the leaves of the same plant or from other species such as Laminaria digitata, Ascophyllum sp, and Lessonia sp. Alginate can also to some extent be fractionated, but the efficiency is low (Haug et al. 1967).
98
H. Ertesvåg et al.
Another advantage of using seaweed alginates compared with using those from bacteria is that they can generally be produced more cheaply from seaweeds than from bacteria. However, bacterial alginates not only present a wider structural variety, but they also offer the potential of designing new alginate structures to meet demands for alginates with improved or new functional properties. For many types of application, such as for pharmaceuticals, price per unit weight is not necessarily the most critical parameter, but rather the product quality in terms of homogeneity and structural reproducibility and availability. For such purposes there is certainly a potential for using alginates produced by bacteria, as one could envision both stable production and unique quality. To achieve this one must be able to tailor-make the alginate structures in vitro or in vivo, and in this chapter the methods available for doing this are discussed.
2
Mannuronan C-5-Epimerases
Homopolysaccharides with only one type of bond are synthesized using a nucleotide sugar and one glycosyltransferase. This concept may be extended to linear heteropolysaccharides by suggesting that several glycosyltransferases cooperate, and that the polysaccharide structure will be a product of their relative reaction rates, substrate requirements, and precursor availability. This was also suggested for the biosynthesis for alginate, as Lin and Hassid (1966b) claimed to be able to isolate the potential precursors GDP-d-mannuronic acid and GDP-l-guluronic acid from the brown algae Fucus gardnerii. The same authors also showed the presence of enzymes necessary to incorporate mannose 1-phosphate into alginate using 14C-labeled mannose derivatives (Lin and Hassid 1966a). However, Haug and Larsen (1969) were able to show that the bacterium A. vinelandii secretes an enzyme that epimerizes M-residues to G-residues on the polymer chain, and this was the first example of an epimerase working on a polymer. This suggested that alginate is synthesized as a homopolymer, as was later confirmed by Chitnis and Ohman (1990). A mannuronan C-5-epimerase was also identified in the algae Pelvetia canaliculata (Magdwick et al. 1973). Mammalian cells were found to synthesize heparin in a similar way, where a glucuronan C-5-epimerase converts some of the d-glucuronic acid residues to l-iduronic acid residues (Lindahl et al. 1972).
2.1
Early Studies on the Secreted Mannuronan C-5-Epimerase from A. vinelandii
Mannuronan C-5-epimerase could be partly purified from the A. vinelandii culture broth using ammonium sulfate precipitation (Larsen and Haug 1971). Further purification was later obtained by using ion-exchange and affinity chromatography (Skjåk-Bræk and Larsen 1982b; Skjåk-Bræk et al. 1985). When the partially
Enzymatic Alginate Modification
99
purified enzyme was incubated with alginate in the presence of tritiated water most of the radioactivity was found in the G residues, but a smaller amount was also found in M residues (Larsen and Haug 1971). The authors suggested a model in which the first step in the catalytic reaction is the abstraction of H-5 followed by addition of hydrogen from water to the other side of the uronic acid residue. The carbazole method can also be used to follow the epimerization reaction (Knutson and Jeanes 1968; Haug and Larsen 1971), since G gives rise to a higher color intensity than M. Infrared light may also give an estimate of the G content (Mackie 1971), and it is possible to measure the increase in G content using G-specific lyases as was first described by Currie and Turvey (1982). However, a more accurate method involves the use 3H-5-labeled mannuronan as a substrate. During epimerization, this hydrogen atom is exchanged with water and the reaction can be measured by quantifying the radioactivity remaining in the soluble phase after the alginate has been precipitated by ethanol (SkjåkBræk and Larsen 1982a). These methods can quantify the increase in G content as the epimerization reaction proceeds, but cannot provide information regarding the distribution pattern of the G residues (G-blocks or MG-blocks). This can to some extent be achieved by partial acid hydrolysis of the polymer followed by acid precipitation, since the MG-blocks are more soluble in acid than G-blocks and M-blocks (Haug et al. 1967). However, 1H-NMR and 13C-NMR spectroscopy (Grasdalen et al. 1979, 1981; Grasdalen 1983) is now the preferred method since it can be used to obtain accurate data on the frequencies of the different dimer and trimer distributions in the alginate polymer. Even in the first papers describing the secreted mannuronan C-5-epimerases it was mentioned that the concentration of divalent calcium ions affected activity (Haug and Larsen 1969, 1971). However, if the reaction was allowed to proceed to the end point, an intermediary concentration of Ca2+ (about 1 mM) gave rise to the highest number of G residues. Ca2+ also increased the stability of the enzyme. Later, with use of purified epimerase, it was confirmed that at low concentrations of Ca2+ G-blocks were preferably made, while at higher concentrations MG-blocks dominated (Larsen et al. 1986; Skjåk-Bræk et al. 1986b). One hypothesis that could explain this observation is that the increase in Ca2+ changed the action pattern of the enzyme from a predominantly single-chain mechanism to a multipleattack mechanism with increasing number of calcium ions (Larsen et al. 1986). However, in a report by Ofstad and Larsen (1981), it was stated that different batches of ammonium sulfate precipitated epimerase varied both regarding their optimal calcium ion concentration for activity and the monomer sequence distribution in the polymer reaction product. These authors suggested that the lack of reproducibility could be explained by assuming the presence of two enzymes with different calcium optimums and preference with respect to the epimerization patterns introduced. The relative amounts of these two enzymes could then vary among different batch cultivations of the bacterium. As discussed later, this prediction later turned out to be correct, except that the complexity is even greater than they anticipated.
100
2.2
H. Ertesvåg et al.
Cloning and Characterization of the Mannuronan C-5-Epimerases from A. vinelandii
To resolve the issues described in the previous section and also to obtain larger quantities of the epimerase, the gene or genes encoding them had to be cloned. The purified proteins were separated by sodium dodecyl sulfate polyacrylamide gel elecrophoresis, and the N-terminal sequence of the protein corresponding to the most intense band (122 kDa) was determined. A degenerate nucleotide probe was constructed on the basis of this information and was used to screen a genomic bacteriophage l library of A. vinelandii (Ertesvåg et al. 1994). Positive clones were identified, and parts of one of these were subcloned into a plasmid vector. Surprisingly, several of the plasmids expressed active mannuronan C-5-epimerases in Escherichia coli (Ertesvåg et al. 1994, 1995). The insert in the phage was then sequenced and it was found to contain a gene encoding a protein with the expected N-terminal sequence, but also genes encoding parts of two other homologous proteins. Further analyses of other positive phage clones finally resulted in the identification of five different homologous genes (Ertesvåg et al. 1995). Furthermore, Southern blotting revealed that the A. vinelandii chromosome contains even more homologous sequences. The last of these genes were finally cloned by a second screen of the genomic library (Svanem et al. 1999), and it is now clear that the total number of homologous genes is eight, but only seven of them (algE1–algE7) encode proteins displaying mannuronan C-5-epimerase activity after expression in E. coli. The product of the last gene, designated “algY,” has no known function (Svanem et al. 1999). The secreted epimerases are all modular enzymes containing one or two A-modules and one to seven R-modules (Fig. 2). Within each module type the range of homology is between 50 and 100%. The two A-modules from AlgE1 were expressed separately in E. coli and both were shown to be active epimerases (Ertesvåg and Valla 1999), proving that this module contained the catalytic site of the enzyme. The presence of the R-modules increased the activity of the enzyme about tenfold. Interestingly, this increase was found when the R-modules were located both N-terminally and C-terminally with respect to the A-module (Ertesvåg and Valla 1999). The R-modules contain four to seven imperfect direct repeats of a nine amino acid long motif found in many proteins exported by ATP binding cassette transporters (Delepelaire 2004). Such type I secreted proteins do not contain an N-terminal secretion signal, but some feature of their C-terminal sequence is necessary for secretion. The genes (eexDEF) needed for secretion of the epimerases have been identified (Gimmestad et al. 2006). The nonameric repeat has been shown to bind calcium (Baumann et al. 1993), and 3D-structure studies of the R-module of AlgE4 support this conclusion (Aachmann et al. 2006). Furthermore, binding and 3D-structure studies of the AlgE4 A-module have demonstrated that this catalytic unit also binds Ca2+ (Ertesvåg and Valla 1999; Rozeboom et al. 2008). The different epimerases produce alginates with distinct epimerization patterns; AlgE2 and AlgE5 seem to predominantly epimerize M residues located next to a
Enzymatic Alginate Modification
101
Modular structure
Enzyme
A
End products
Other activities
G-and MGblocks
AlgE1
A
R
R
R
R
AlgE2
A
R
R
R
AlgE3
A
R
R
R
AlgE4
A
R
AlgE5
A
R
R
R
G-blocks (medium)
AlgE6
A
R
R
R
G-blocks (long)
AlgE7
A
R
R
R
G-and MGblocks
Lyase activity
PsmE
A
R
R M R
G-blocks (long)
Deacetylase
G-blocks (short)
R A
R
R
R R
Weak lyase activity
G-and MGblocks MG-blocks
N
RTX RTX
Weak lyase activity
Fig. 2 Modular structure and end products of the secreted mannuronan C-5-epimerases. The relationships within the group of A-modules and R-modules (Svanem et al. 1999) are shown by similarities and differences in cross-hatchings and gray tones. The Pseudomonas syringae enzyme also contains two additional calcium-binding modules (M and RTX) and a module displaying alginate deacetylase activity (N). The weak lyase activity of AlgE2 and AlgE5 (0.1–0.3% of the reactions) is probably a result of failed epimerase reactions
preexisting G residue, generating G-blocks (Ertesvåg et al. 1999; Ramstad et al. 1999), while AlgE4 does not do this, and therefore forms alternating sequences, (MG)n, in the reaction product (Høidal et al. 1999). As can be seen from Fig. 2, the A-modules of AlgE2 and AlgE5 belong to one homology group of A-modules, while the A-module of AlgE4 belongs to another group. AlgE1 and AlgE3 both contain two A-modules, one from each of these two groups. And when they were expressed separately, the A1-modules make G-blocks, while the A2-modules introduce alternating structures (Ertesvåg et al. 1998b, 1999). However, the A-module of AlgE6 groups with that of AlgE4, but still this enzyme introduces G-blocks into the alginate (Svanem et al. 1999). A study using hybrid A-modules between AlgE2A and AlgE4A might reconcile these findings, since it was shown that only certain parts of the A-module are important for the epimerization pattern (Bjerkan et al. 2004). The A-module of AlgE7 is similar to that of AlgE6. However, the end product after incubation of mannuronan with AlgE7 is relatively small oligomers with G residues at the reducing end and an unsaturated residue at the nonreducing end, identifying AlgE7 as being both a mannuronan C-5-epimerase and an alginate lyase (Svanem et al. 2001). The same residues are important for both the epimerase and lyase activities, emphasizing the hypothesis proposed by Gacesa (1987) that alginate epimerases and lyases share the first steps of the catalysis.
102
H. Ertesvåg et al.
The grouping of the epimerases into those generating G-blocks and those preferably introducing alternating structures is fairly crude; there are differences both in the relative numbers of G-blocks and alternating structures made (Fig. 2), the acceptance of alternating structures as a substrate for further epimerization, and the requirements for Ca2+ (Ertesvåg et al. 1999; Campa et al. 2004; Holtan et al. 2006). In the more biological context it has been shown that different epimerases are expressed at different times in the life cycle of A. vinelandii, probably reflecting a need for different alginates in the vegetative capsule and in the cyst (Høidal et al. 2000). The crucial role of these enzymes in cyst formation was recently confirmed by studies of the phenotypes of mutants in which all the algE genes had been deleted (Steigedal et al. 2008). A gene encoding a secreted mannuronan C-5-epimerase designated PsmE has also been identified in P. syringae (Bjerkan et al. 2004). This enzyme contains one A-module and three R-modules (Fig. 2). A second module putatively binding calcium is interspersed between the last two R-modules. These modules are followed by a module encoding an alginate deacetylase and lastly by a third type of a calcium-binding module. PsmE was produced recombinantly in E. coli and was shown to introduce G-blocks into alginate. Recombinantly produced PsmE is able to remove acetyl groups on acetylated alginate and thus uses acetylated alginate as a substrate for epimerization, in contrast to the AlgE epimerases (Bjerkan et al. 2004). G-block-containing alginates have not been reported from Pseudomonas sp. However, psmE was identified as a gene expressed at low (18°C) temperature (Ullrich et al. 2000), and it is possible that alginate has not yet been prepared from the relevant strains growing at the conditions required for production of such an alginate. A-modules are only found in the genomes of A. vinelandii and P. syringae among all the genomes presently deposited in GenBank.
2.3
Cloning and Characterization of the Bacterial Periplasmic Mannuronan C-5-Epimerases
Chitnis and Ohman (1990) were able to isolate two mutants of P. aeruginosa that produced alginate with no G residues, and cloned a DNA fragment that complemented this defect. The DNA fragment was later expressed in E. coli and was shown to encode a mannuronan C-5-epimerase that was localized in the periplasm (Franklin et al. 1994). The enzyme was inhibited by acetyl groups on the alginate. Later, similar enzymes were cloned and characterized from A. vinelandii (Rehm et al. 1996) and P. fluorescens (Morea et al. 2001; Gimmestad et al. 2003). All three characterized AlgG epimerases introduce only single G residues (no G-blocks) into the alginate. However, in a later study, Jerga et al. (2006) claimed that AlgG from P. aeruginosa can produce G-blocks in vitro. AlgG does not need calcium for activity (Rehm et al. 1996; Jerga et al. 2006). AlgG is not only needed for epimerization of the nascent polymer, it also has a structural function in transporting the polymer out of the cell, presumably by being
Enzymatic Alginate Modification
103
part of a periplasmic protein complex (Gimmestad et al. 2003; Jain et al. 2003). The A. vinelandii AlgG epimerase has fairly low activity in vivo and in vitro (Rehm et al. 1996; Steigedal et al. 2008), indicating that the AlgE epimerases determine the G distribution in this bacterium.
2.4
Mannuronan C-5-Epimerases from Algae
As already mentioned, mannuronan C-5-epimerase activity has been described from several brown algae. However, genes encoding putative algal epimerases seem to be difficult to express recombinantly. Nyvall et al. (2003) were able to show that L. digitata expresses several genes sharing significant sequence similarities with the bacterial epimerases, especially the AlgG group. The expression pattern of these genes coincided with the level of mannuronan C-5-epimerase, strengthening the hypothesis that these genes encode such enzymes. The data indicated that L. digitata encodes a family of at least 21 mannuronan C-5-epimerases. It seems probable that the high number of different epimerases enables the algae to tailor alginates according to the needs of the season and tissue concerned, an explanation analogous to that for the existence of the AlgE epimerase family in A. vinelandii. Alginate-related genes from red algae have not yet been reported.
2.5
Structure of the Mannuronan C-5-Epimerases
The 3D structures of the A- and R-modules of AlgE4 were recently published (Aachmann et al. 2006; Rozeboom et al. 2008), and were both found to be right-handed parallel b-rolls (Fig. 3a). The A-module consists of one b-helix with positively charged residues facing a shallow groove. One calcium ion was identified in the structure (Fig. 3b) (Rozeboom et al. 2008). When the crystal was soaked in a solution containing a tetramer of M residues, it was found that the oligomer bound with the reducing end closest to the N-terminal end of the A-module. The R-module forms a narrower b-roll that also binds an oligomer of M residues. On the basis of this, it was proposed that the whole enzyme probably forms one elongated protein with the positively charged alginate binding site spanning both modules (Aachmann et al. 2006). The catalytic site of AlgE4 seems to consist of Asp152, His154, and Tyr149 (Fig. 3b). This assumption is consistent with the structure data, and even conservative substitutions of one of these amino acids decreased the enzyme activity to less than 1% (Rozeboom et al. 2008). The studies also showed that this catalytic site is very similar to the corresponding site reported for five alginate lyases. The similarity between the lyases was first observed by Osawa et al. (2005). AlgG from P. aeruginosa is predicted to be structurally similar to that of the A-module (Douthit et al. 2005), and also shares some sequence similarity to AlgE4 around the catalytic site.
104
H. Ertesvåg et al.
a
A-module
R-module
b
Fig. 3a Electrostatic surface model of the A- and R-modules of AlgE4. b The catalytic site of AlgE4 is enlarged and shown in stick representation. The bound Ca2+ ion is shown as an orange sphere. The models were constructed using MOLMOL (Koradi et al. 1996) and PyMOL (DeLano 2002). (a Reproduced with permission from Achmann et al. 2006)
Rozeboom et al. (2008) also proposed a catalytic mechanism for the epimerases based on the structure and activities of mutant proteins (Fig. 4). In this model Asp152 (or possibly Asp178) protonates the +1 M residue. Tyr149 abstracts the C-5 proton from the protonated sugar, while His154 subsequently donates a proton to C-5 from the other side of the sugar.
2.6
Mode of Action of the Secreted Mannuronan C-5-Epimerases
Each of the AlgE1–AlgE7 epimerases generates specific nonrandom epimerization patterns when acting upon mannuronan or any other alginate as a substrate. Such nonrandom block structures are typical in all alginates and have traditionally been attributed to the processive mode of action of polymer-modifying enzymes. “Processivity” refers to the average number of times a reaction is repeated between association and dissociation of the enzyme-substrate complex. High processivity is common for enzymes taking part in the replication and modification of RNA and DNA (Kornberg and Baker 1992). However, any enzyme that has more than one substrate-binding subsite and performs multiple modifications on a substrate may display processivity (Hartmann et al. 2002). Several examples of this type of action have been confirmed for polysaccharides, mainly for exolytic enzymes (Ernst et al. 1998;
Enzymatic Alginate Modification
a
152
b
149
O e O H N 154
O H
152
−OOC H
H RO N
178
d
O H
152 O e O
+ H O
N
154
OH
O
O H
H+
OH R1 COO−
H O
e 178 O
e O
O H H O eO O H O RO N OH OH + N 154 H O e 178
O
OH
OH
+
e
149
O e O
c
149
Alginate
N
105
H
OH COO−
154
OH
H+ OH
O H
f
R1 COO−
149
O H RO HO
O
HO
:N
R1
O H H O H O O H RO O OH N OH + N 154 H O e O 178
152 O e O
149
152 O e O + H
OH O
149
N
H COO− OH O O OH H
H O 178
R1
COO− 154
e O
O H
152 O e O H+ H N Alginate + N
H O e O 178
Fig. 4 Proposed catalytic mechanism of AlgE4. The alginate enters the catalytic site (a, b). The carboxylate group is protonated, allowing it to form a hydrogen bond with Asp152 (or Asp178) (c). Tyr149 is deprotonated (via Arg195), and the alkoxide group extracts H-5 from the mannuronic acid residue in the catalytic site (c). A double bond is formed, resulting in a nearly planar hex-4enepyranosyluronate residue (d). The protonated His145 performs a nucleophilic attack on the C-5 atom of this sugar from the si face, flipping the sugar ring into the 1C4 chair conformation of guluronic acid (d, e). The carboxylic acid moiety is deprotonated and leaves the active site (e, f). His154 is protonated again and the epimerase is ready to perform a new reaction (f). (Modified with permission from Rozeboom et al. 2008)
Breyer and Matthews 2001). In alginates, enzyme processivity influences whether G units are introduced at random or successively as blocks. The formation of long blocks could thus be explained by a high degree of processivity. However, these nonrandom patterns can also arise from a preferred attack mechanism by which the affinity for the substrates depends on preexisting G residues. To distinguish between these two mode of action experimentally is, however, difficult. Results from kinetic experiments using time resolved NMR spectroscopy and modeling based on Monte Carlo simulation suggest that AlgE4 acts processively (Høidal et al. 1999; Hartmann et al. 2002). Evidence for processivity also comes from single molecular pair unbinding studies of mannuronan and AlgE4 using atomic force microscopy (Sletmoen et al. 2004). The degree of processivity was estimated, using specific degrading enzymes, NMR, electrospray ionization mass spectrometry, and capillary electrophoresis. On average, AlgE4 epimerizes ten M residues before the enzyme-substrate complex dissociates, implying that the enzyme, converting every second residue, propagates a distance corresponding to 20 M residues (Campa et al. 2004).
106
H. Ertesvåg et al.
AlgE1 and AlgE6 introduce G-blocks in a processive manner when acting on polyMG as a substrate. (Holtan et al. 2006). The remaining epimerases act according to either a preferred attack mechanism (AlgE2 and AlgE5) or a combination of the two mechanisms. A polymer-modifying enzyme, independent of whether it acts processively or in a preferred mode, would have to interact with more than one residue in the polymer substrate. AlgE4 requires a hexameric mannuronan oligomer as the minimum size of the substrate to support epimerase activity, while seven to eight residues and eight to ten residues appear to be the minimum size for AlgE6 and AlgE1, respectively (Holtan et al. 2006). The recently elucidated crystal structure (Rozeboom et al. 2008) of the A-module reveals that AlgE4A has the structural basis for processivity, since the substrate is partially enclosed in a groove with a relatively large interaction surface, and with a clamp at each side of the active site that would enclose the polymer and help it to slide rather than dissociate.
2.7
Enzyme Specificity; Epimerization of Nonnatural Uronans and Poly(guluronic acid)
The AlgE epimerases are specific for polymeric b-1• 4-linked mannuronic acid found naturally in alginates or in C-6-oxidized mannose-containing polysaccharides. AlgE1, AlgE4, and AlgE6 are able to epimerize M-residues in various nonnatural uronans obtained by a selective 2,2,6,6-tetramethylpiperidine-1-oxyl-mediated hypohalite oxidation of mannose containing polysaccharides, leading to new polyelectrolytes with enhanced gelling properties in aqueous systems containing Ca2+ ions (Crescenzi et al. 2000, 2002). The epimerization reaction of d-ManA to l-GulA is analogous to the conversion of d-GlcA into l-IdoA (Lindahl et al. 1972; Hagner-McWhirter et al. 2000) in the biosynthesis of heparin, heparan sulfate, and dermatan sulfate, and glucuronan C-5epimerase is the only other known polymer-level C-5-epimerase. It has therefore been of considerable interest to establish if mannuronan C-5-epimerases also could epimerize polymer-linked GlcA. In 2000 an astonishing paper by Chang et al. (2000) was published in which they claimed that epimerase isolated from A. vinelandii was able to epimerize polymer-linked uronic acid in a range of C-6-oxidized polymers (cellulose, amylose, pullulan, and chitosan) and in native pectin. Over the years, we and other groups have tried to reproduce their results but without success (Holtan 2006) (A. Hotchkiss, USDA ARS, USA, and V. Crescenzi and M. Dentini, University La Sapienza, Rome, Italy, personal communications). Both wild-type and recombinant epimerases have been tested alone or in combinations on the following substrates: C-6-oxidized cellulose and amylose, microbially produced glucuronan from Rhizobium meliloti, pectate, and heparin precursors such as E. coli K5 capsular polysaccharide and sulfaminoheparosan. No epimerization was detected. There is, however, indicative evidence that the epimerase binds to GlcA-containing polymers. In a study on the action of the mannuronan C-5-epimerases on C-6-oxidized
Enzymatic Alginate Modification
107
konjac glucomannan (Crescenzi et al. 2002) a GlcA-a-l-(1→4)GulA-b-d-(1→4) ManA sequence could be identified after incubation with AlgE4 and AlgE6. This indicates that the AlgE4 subsite (−1) is able to bind to b-d-glucuronic acid. Increasing the AlgE4 and AlgE6 incubation time led to depolymerization of the polymer by a b-elimination mechanism. Similar observations have been reported for C-6-oxidized cellulose (b-(1→4)-d-glucuronic acid) after incubation with AlgE4 (Crescenzi et al. 2000; Crescenzi et al. 2002). This indicates that GlcA, which is a C-2-epimer of ManA, can to some extent be accepted in the active site of AlgE4 mannuronan C-5-epimerase such that the proton on C-5 is abstracted although epimerization to l-iduronic acid is not taking place. Because GlcA is able to recognize b-(1,4)-d-glucuronic acid and act as a lyase, only subtle conformational modifications of AlgE4 or another epimerase might turn it into an active glucuronan C-5-epimerase.
2.8
In Vitro Epimerization with AlgE Epimerases
Alginate gels represent very challenging semisolid structures with applications in biotechnology and industrial fields. The ability to control and tune the mechanical properties represents a key feature for improving the applications of such biomaterial. The availability of AlgE epimerases provides a powerful tool to design the sequential structures and allows monitoring of the effect of compositional modifications on the functional properties of alginate gels. The AlgE enzymes, which in their natural state act in the exocellular environment, are soluble, stable, and require no cofactor except calcium ions. They have been proven to be very effective in modifying all types of alginate in vitro (Mørch et al. 2007). In particular AlgE1, AlgE4, and AlgE6 are technically useful owing to their lack of lyase activity. AlgE4 is highly effective in converting M-blocks into polyMG sequences in any block-structured alginate. When acting on pure mannuronan, this enzyme is able to convert the entire polymer into polyMG (Fig. 1) without breaking the polymer backbone (Hartmann et al. 2002). As described earlier, AlgE1 and AlgE6 generate long G-blocks either by elongating preexisting G-blocks or by condensing them by epimerizing single M residues flanked by Gs (Fig. 1). PolyG is still difficult to make since the calcium ions, required by the enzyme, are gradually depleted owing to their cooperative binding to the G-blocks generated. Moreover, binding of calcium ions could lead to gel formation, rendering the molecules less accessible as a substrate for the enzyme. With use of suboptimal concentrations of calcium ions in the presence of high concentrations of antigelling ions such as sodium, high molecular weight alginates with 97% G were obtained (Holtan et al. 2006). Since three to four M residues at the reducing and nonreducing ends cannot be attacked by the enzyme, this probably represents the end point of the epimerization. The availability of pure alginates resembling separately the three extreme block sequences present in the natural polysaccharide, namely, G-blocks, M-blocks, and MG-blocks, was found to be a fundamental tool to point out their role in the final hydrogel (Donati et al. 2005).
108
3
H. Ertesvåg et al.
Alginate Lyases
The viscosity of alginates is mostly dependent on molecular weight, although the G distribution affects the stiffness of the chain and thereby also the viscosity. Alginate lyases split the polymer by b-elimination, leaving an unchanged saturated uronate on the reducing end and an unsaturated (4-deoxy-l-erythro-hex-4enepyranosyluronate) residue, symbolized by D, at the nonreducing end. Alginate lyases are widely distributed in nature, in organisms growing on alginate as a carbon source, in alginate-producing organisms, and in some bacteriophages (Hashimoto et al. 2009). In this review we concentrate on lyases produced by alginate-producing bacteria and those used to characterize alginate structure.
3.1
Alginate Lyases as Tools To Determine Alginate Structure
Specific degrading enzymes are invaluable tools for structural analysis of macromolecules. This is also the case with alginate-specific degrading enzymes, of which all known can be classified as lyases. The specificity of these enzymes is commonly directed towards one of the residues in the glycosidic linkage, i.e., the lyase from Klebsiella aerogenes cleaves G-G and G-M specifically for the guluronate in the glycon position, leaving a reducing guluronate end and an unsaturated end originating from either M or G (Boyd and Turvey 1978). The lyase from A vinelandii is specific for M in the glycon position, cleaving M-M and M-G bonds (Ertesvåg et al. 1998a). Other enzymes are specific for the sugars on both sides of the bond, such as the M-M-specific lyase from P. alginovora (Chavagnat et al. 1996). Although the chemical identity of the aglycon is lost by the lyase-catalyzed reaction, the unsaturated end functions as a chromophore (UV 230 nm) and is also easily identified by NMR. Combined with mass spectrometry and NMR, specific lyases have been used successfully in analyses of subsite specificity for the epimerases (Campa et al. 2004) as well as in analysis of block length and block length distribution (Holtan et al. 2006). We are currently using enzyme engineering in our search for more specific lyases to be used in alginate sequencing.
3.2
An Alginate Lyase Is Needed To Remove Aberrantly Localized Alginate
In all alginate biosynthetic gene clusters sequenced so far a gene, algL, encoding an alginate lyase has been found. Initially this was surprising since degradation of a newly synthesized polymer seems a futile reaction. It was suggested that AlgL could be involved in determination of the polymer length, making primers for alginate synthesis or releasing the cells from alginate (Boyd et al. 1993). Later it
Enzymatic Alginate Modification
109
was shown that AlgG, AlgK, and probably AlgX form a protein complex guiding the nascent alginate chain from the polymerase in the inner membrane and through the secretion pore AlgE (AlgJ in A. vinelandii) in the outer membrane (Rehm et al. 1994; Rehm 1996; Jain and Ohman 1998; Gimmestad et al. 2003; Jain et al. 2003; Robles-Price et al. 2004). If any of these three proteins are missing, the product is unsaturated oligomers, suggesting that the protein complex protects the alginate against AlgL. It was then shown that alginate production in the absence of AlgL is lethal to the cells (Bakkevig et al. 2005; Jain and Ohman 2005). The explanation seems to be that the polyanion alginate in the periplasm will attract cations and thus create an osmotic stress that leads to cell lysis. If one of the alginate-producing complexes does not function properly, this will happen, and AlgL is present in the periplasm as a repair system to remove these lethal molecules. P. aeruginosa has been found to encode a second alginate lyase (Yamasaki et al. 2004) that prefers MG-rich alginate. The biological function of this lyase is not known.
4
Alginate Acetylation
Bacterial, but not algal, alginates are acetylated at O2, O3, or both (Skjåk-Bræk et al. 1986a). Acetylation increases the water-binding capacity of alginate and its viscosity (Skjåk-Bræk et al. 1989). It has also been shown that acetylation increases intermolecular association of alginate molecules (Windhues and Borchard 2003). The genes, algI, algJ (algV in A. vinelandii), and algF encoding the proteins needed for acetylation are encoded in the alginate biosynthetic gene clusters (Franklin and Ohman 1996; Vazquez et al. 1999). However, the gene products are not necessary for alginate production (Shinabarger 1993). Still, acetylation takes place in the periplasm (Franklin and Ohman 2002), indicating that although the nascent alginate is not accessible to AlgL it is accessible to AlgF. For P. aeruginosa, acetylation is necessary for the formation of microcolonies during the early stages of biofilm development (Tielen et al. 2005). This suggests that acetylation of alginate is important for cell–cell attachment. Acetylation will also protect the alginate against degradation by many alginate lyases. For A. vinelandii, the G distribution of the alginate is determined by the secreted mannuronan C-5epimerases. Since these enzymes will not epimerize acetylated residues, acetylation may be a way for the bacterium to determine the final degree of epimerization.
5
Use of Alginate-Modifying Enzymes for Biotechnological Purposes
The studies on alginate biosynthesis have established that alginate is synthesized as the homopolymer mannuronan, and that the diversity of structural and functional properties found in different alginates results from the action of C-5-epimerases,
110
H. Ertesvåg et al.
lyases, and acetylases. So far the secreted mannuronan C-5-epimerases have the greatest potential from a biotechnological point of view. Pure mannuronan can now be produced by mutants lacking epimerases (Gimmestad et al. 2003). When this polymer is used as a substrate for the AlgE epimerases, alone or in combination, it is thus for the first time feasible to produce alginates with predetermined structures. In addition to compositionally homogeneous alginates (mannuronan, guluronan, and polymers with strictly alternating M–G) (see Fig. 1), this could also include other alginates with extreme composition and sequential structure not found in nature. Of particular interest is a type of alginate comprising only G-blocks interspaced with polyMG sequences. This material, which was made by a two-step conversion of polyM with AlgE4 and AlgE1, forms very strong, elastic, nonswelling gels, particularly suitable for making immune-protective microcapsules for cell transplantations (Mørch et al. 2007, 2008). Specific alginate lyases have also been used for preparation of pure uronate blocks (Heyraud et al. 1998). Studies on oligomeric G-blocks have shown that they can affect many properties of an alginate gelling system (Jørgensen et al. 2007). Alginate lyases have also been suggested as mucolytic agents in the treatment of cystic fibrosis (Mrsny et al. 1994) A further challenge would be to produce tailored alginates in vivo. This might be done by first making an epimerase-negative strain (including an epimerization-defective mutation in algG) of A. vinelandii and then introducing specific epimerase genes controlled by a heterologous promoter. A. vinelandii has a high alginate lyase activity which needs to be controlled to obtain alginate of a sufficiently high molecular weight, but it has the ability to secrete the AlgE epimerases and has been shown to make alginates with a high fractional content of G residues (Skjåk-Bræk et al. 1986a). Acknowledgements The authors want to thank Wenche Iren Strand and Finn L. Aachmann for preparing some of the figures.
References Aachmann FL, Svanem BI, Güntert P, Petersen SB, Valla S, Wimmer R (2006) NMR structure of the R-module: a parallel beta-roll subunit from an Azotobacter vinelandii mannuronan C-5 epimerase. J Biol Chem 281:7350–7356 Bakkevig K, Sletta H, Gimmestad M, Aune R, Ertesvåg H, Degnes K, Christensen BE, Ellingsen TE, Valla S (2005) Role of the Pseudomonas fluorescens alginate lyase (AlgL) in clearing the periplasm of alginates not exported to the extracellular environment. J Bacteriol 187:8375–8384 Baumann U, Wu S, Flaherty KM, McKay DB (1993) Three-dimensional structure of the alkaline protease of Pseudomonas aeruginosa: a two-domain protein with a calcium binding parallel beta roll motif. EMBO J 12:3357–3364 Bilan MI, Usov AI (2001) Polysaccharides of calcareous algae and their effect on the calcification process. Russ J Bioorg Chem 27:2–16 Bjerkan TM, Bender CL, Ertesvåg H, Drabløs F, Fakhr MK, Preston LA, Skjåk-Bræk G, Valla S (2004) The Pseudomonas syringae genome encodes a combined mannuronan C-5-epimerase
Enzymatic Alginate Modification
111
and O-acetyl hydrolase, which strongly enhances the predicted gel-forming properties of alginates. J Biol Chem 279:28920–28929 Boyd J, Turvey JR (1978) Structural studies of alginic acid using a bacterial poly-a-L-guluronate lyase. Carbohydr Res 66:187–194 Boyd A, Ghosh M, May TB, Shinabarger D, Keogh R, Chakrabarty AM (1993) Sequence of the algL gene of Pseudomonas aeruginosa and purification of its alginate lyase product. Gene 131:1–8 Breyer WA, Matthews BW (2001) A structural basis for processivity. Protein Sci 10:1699–1711 Campa C, Holtan S, Nilsen N, Bjerkan TM, Stokke BT, Skjåk-Bræk G (2004) Biochemical analysis of the processive mechanism for epimerisation of alginate by mannuronan C-5 epimerase AlgE4. Biochem J 381:155–164 Chang PS, Mukerjea R, Fulton DB, Robyt JF (2000) Action of Azotobacter vinelandii poly-betaD-mannuronic acid C-5-epimerase on synthetic D-glucuronans. Carbohydr Res 329:913–922 Chavagnat F, Duez C, Guinand M, Potin P, Barbeyron T, Henrissat B, Wallach J, Ghuysen JM (1996) Cloning, sequencing and overexpression in Escherichia coli of the alginate lyase-encoding aly gene of Pseudomonas alginovora: identification of three classes of alginate lyases. Biochem J 319:575–583 Chitnis CE, Ohman DE (1990) Cloning of Pseudomonas aeruginosa algG, which controls alginate structure. J Bacteriol 172:2894–2900 Crescenzi V, Hartmann M, de Nooy AE, Rori V, Masci G, Skjåk-Bræk G (2000) Epimerization of nonnatural uronans with mannuronan C-5-epimerases to obtain alginatelike polysaccharides. Biomacromolecules 1:360–364 Crescenzi V, Skjåk-Bræk G, Dentini M, Masci G, Bernalda MS, Risica D, Capitani D, Mannina L, Segre AL (2002) A high field NMR study of the products ensuing from konjak glucomannan C(6)-oxidation followed by enzymatic C(5)-epimerization. Biomacromolecules 3:1343–1352 Currie AJ, Turvey JR (1982) An enzymatic method for the assay of D-mannuronan C-5 epimerase activity. Carbohydr Res 107:159–161 DeLano WL (2002) The PyMOL molecular graphics system, vol 228. DeLano Scientific, Palo Alto Delepelaire P (2004) Type I secretion in Gram-negative bacteria. Biochim Biophys Acta 1694:149–161 Donati I, Holtan S, Mørch YA, Borgogna M, Dentini M, Skjåk-Bræk G (2005) New hypothesis on the role of alternating sequences in calcium-alginate gels. Biomacromolecules 6:1031–1040 Douthit SA, Dlakic M, Ohman DE, Franklin MJ (2005) Epimerase active domain of Pseudomonas aeruginosa AlgG, a protein that contains a right-handed beta-helix. J Bacteriol 187:4573–4583 Draget K, Skjåk-Bræk G, Smidsrød O (1994) Alginic acid gels. The effect of alginate chemical composition and molecular weight. Carbohydr Polym 25:31–38 Draget KI, Strand B, Hartmann M, Valla S, Smidsrød O, Skjåk-Bræk G (2000) Ionic and acid gel formation of epimerised alginates; the effect of AlgE4. Int J Biol Macromol 27:117–122 Ernst S, Rhomberg AJ, Biemann K, Sasisekharan R (1998) Direct evidence for a predominantly exolytic processive mechanism for depolymerization of heparin-like glycosaminoglycans by heparinase I. Proc Natl Acad Sci U S A 95:4182–4187 Ertesvåg H, Valla S (1999) The A modules of the Azotobacter vinelandii mannuronan-C-5-epimerase AlgE1 are sufficient for both epimerization and binding of Ca2+. J Bacteriol 181:3033–3038 Ertesvåg H, Doseth B, Larsen B, Skjåk-Bræk G, Valla S (1994) Cloning and expression of an Azotobacter vinelandii mannuronan C-5-epimerase gene. J Bacteriol 176:2846–2853 Ertesvåg H, Høidal HK, Hals IK, Rian A, Doseth B, Valla S (1995) A family of modular type mannuronan C-5-epimerase genes controls alginate structure in Azotobacter vinelandii. Mol Microbiol 16:719–731 Ertesvåg H, Valla S, Skjåk-Bræk G (1996) Genetics and biosynthesis of alginates. Carbohydr Eur 14:14–18 Ertesvåg H, Erlien F, Skjåk-Bræk G, Rehm BH, Valla S (1998a) Biochemical properties and substrate specificities of a recombinantly produced Azotobacter vinelandii alginate lyase. J Bacteriol 180:3779–3784
112
H. Ertesvåg et al.
Ertesvåg H, Høidal HK, Skjåk-Bræk G, Valla S (1998b) The Azotobacter vinelandii mannuronan C-5-epimerase AlgE1 consists of two separate catalytic domains. J Biol Chem 273: 30927–30932 Ertesvåg H, Høidal HK, Schjerven H, Svanem BI, Valla S (1999) Mannuronan C-5-epimerases and their application for in vitro and in vivo design of new alginates useful in biotechnology. Metab Eng 1:262–269 Flo TH, Ryan L, Kilaas L, Skjåk-Bræk G, Ingalls RR, Sundan A, Golenbock DT, Espevik T (2000) Involvement of CD14 and beta2-integrins in activating cells with soluble and particulate lipopolysaccharides and mannuronic acid polymers. Infect Immun 68:6770–6776 Franklin MJ, Ohman DE (1996) Identification of algI and algJ in the Pseudomonas aeruginosa alginate biosynthetic gene cluster which are required for alginate O acetylation. J Bacteriol 178:2186–2195 Franklin MJ, Ohman DE (2002) Mutant analysis and cellular localization of the AlgI, AlgJ, and AlgF proteins required for O acetylation of alginate in Pseudomonas aeruginosa. J Bacteriol 184:3000–3007 Franklin MJ, Chitnis CE, Gacesa P, Sonesson A, White DC, Ohman DE (1994) Pseudomonas aeruginosa AlgG is a polymer level alginate C5-mannuronan epimerase. J Bacteriol 176:1821–1830 Gacesa P (1987) Alginate-modifying enzymes. A proposed unified mechanism of action for the lyases and epimerases. FEBS Lett 212:199–202 Gimmestad M, Sletta H, Ertesvåg H, Bakkevig K, Jain S, Suh S-j, Skjåk-Bræk G, Ellingsen TE, Ohman DE, Valla S (2003) The Pseudomonas fluorescens AlgG protein, but not its mannuronan C5-epimerase activity, is needed for alginate polymer formation. J Bacteriol 185:3515–3523 Gimmestad M, Steigedal M, Ertesvåg H, Moreno S, Christensen BE, Espín G, Valla S (2006) Identification and characterization of an Azotobacter vinelandii Type I secretion system responsible for export of the AlgE-type mannuronan C5-epimerases. J Bacteriol 188:5551–5560 Gorin PAJ, Spencer JFT (1966) Exocellular alginic acid from Azotobacter vinelandii. Can J Chem 44:993–998 Govan JR, Fyfe JA, Jarman TR (1981) Isolation of alginate-producing mutants of Pseudomonas fluorescens, Pseudomonas putida and Pseudomonas mendocina. J Gen Microbiol 125:217–220 Grasdalen H (1983) High-field, 1-H-n.m.r. spectroscopy of alginate. Sequential structure and linkage-conformation. Carbohydr Res 118:255–260 Grasdalen H, Larsen B, Smidsrød O (1979) A p.m.r. study of the composition and sequence of uronate residues in alginate. Carbohydr Res 68:23–31 Grasdalen H, Larsen B, Smidsrød O (1981) 13C-n.m.r. studies of monomeric composition and sequence in alginate. Carbohydr Res 89:179–191 Hagner-McWhirter A, Lindahl U, Li J (2000) Biosynthesis of heparin/heparan sulphate: mechanism of epimerization of glucuronyl C-5. Biochem J 347:69–75 Hartmann M, Holm OB, Johansen GAB, Skjåk-Bræk G, Stokke BT (2002) Mode of action of recombinant Azotobacter vinelandii mannuronan C-5 epimerases AlgE2 and AlgE4. Biopolymers 63:77–88 Hashimoto W, Maruyama Y, Itoh T, Mikami B, Murata K (2009) Bacterial system for alginate uptake and degradation. Microbiol Monograph (this volume) Haug A, Larsen B (1969) Biosynthesis of alginate. Epimerisation of D-mannuronic to L-guluronic acid residues in the polymer chain. Biochim Biophys Acta 192:557–559 Haug A, Larsen B (1971) Biosynthesis of alginate. II. Polymannuronic acid C-5-epimerase from Azotobacter vinelandii (Lipman). Carbohydr Res 17:297–308 Haug A, Larsen B, Smidsrød O (1967) Studies on the sequence of uronic acid residues in alginic acid. Acta Chem Scand 21:691–704 Haug A, Larsen B, Smidsrød O (1974) Uronic acid sequence in alginate from different sources. Carbohydr Res 32:217–225 Heyraud A, Colin-Morel P, Gey C, Chavagnat F, Guinand M, Wallach J (1998) An enzymatic method for preparation of homopolymannuronate blocks and strictly alternating sequences of mannuronic and guluronic units. Carbohydr Res 308:417–422
Enzymatic Alginate Modification
113
Holtan S (2006) Structural tailoring of alginates. A study of the mode of action of mannuronan C-5-epimerases. Dissertation 2006:117, Norwegian University of Science and Technology, Trondheim Holtan S, Bruheim P, Skjåk-Bræk S (2006) Mode of action and subsite studies of the guluronan block-forming mannuronan C-5 epimerases AlgE1 and AlgE6. Biochem J 395:319–329 Høidal HK, Ertesvåg H, Skjåk-Bræk G, Stokke BT, Valla S (1999) The recombinant Azotobacter vinelandii mannuronan C-5-epimerase AlgE4 epimerizes alginate by a nonrandom attack mechanism. J Biol Chem 274:12316–12322 Høidal HK, Svanem BIG, Gimmestad M, Valla S (2000) Mannuronan C-5 epimerases and cellular differentiation of Azotobacter vinelandii. Environ Microbiol 2:27–38 Jain S, Ohman DE (1998) Deletion of algK in mucoid Pseudomonas aeruginosa blocks alginate polymer formation and results in uronic acid secretion. J Bacteriol 180:634–641 Jain S, Ohman DE (2005) Role of an alginate lyase for alginate transport in mucoid Pseudomonas aeruginosa. Infect Immun 73:6429–6436 Jain S, Franklin MJ, Ertesvåg H, Valla S, Ohman DE (2003) The dual roles of AlgG in C-5epimerization and secretion of alginate polymers in Pseudomonas aeruginosa. Mol Microbiol 47:1123–1133 Jerga A, Raychaudhuri A, Tipton PA (2006) Pseudomonas aeruginosa C5-mannuronan epimerase: steady-state kinetics and characterization of the product. Biochemistry 45:552–560 Jørgensen TE, Sletmoen M, Draget KI, Stokke BT (2007) Influence of oligoguluronates on alginate gelation, kinetics, and polymer organization. Biomacromolecules 8:2388–2397 Knutson CA, Jeanes A (1968) Determination of the composition of the uronic acid mixtures. Anal Biochem 24:482–490 Koradi R, Billeter M, Wuthrich K (1996) MOLMOL: a program for display and analysis of macromolecular structures. J Mol Graph 14:51–55, 29–32 Kornberg A, Baker T (1992) DNA replication, 2nd edn. Freeman, New York Larsen B, Haug A (1971) Biosynthesis of alginate. III. Tritium incorporation with polymannuronic acid 5-epimerase from Azotobacter vinelandii. Carbohydr Res 20:225–232 Larsen B, Skjåk-Bræk G, Painter T (1986) Action pattern of mannuronan C-5-epimerase: generation of block-copolymeric structures in alginates by a multiple-attack mechanism. Carbohydr Res 146:342–345 Lin T-Y, Hassid WZ (1966a) Pathway of alginic acid synthesis in the marine brown alga, Fucus gardneri Silva. J Biol Chem 241:5284–5297 Lin T-Y, Hassid WZ (1966b) Isolation og guanosine diphosphare uronic acids from a marine brown alga, Fucus gardneri Silva. J Biol Chem 241:3282–3293 Lindahl U, Bäckström G, Malmström A, Fransson LA (1972) Biosynthesis of L-iduronic acid in heparin: epimerization of D-glucuronic acid on the polymer level. Biochem Biophys Res Commun 46:985–991 Linker A, Jones RS (1966) A new polysaccharide resembling alginic acid isolated from Pseudomonas. J Biol Chem 241:3845–3851 Mackie W (1971) Semi-quantitative estimation of the composition of alginates by infra-red spectroscopy. Carbohydr Res 20:413–415 Magdwick J, Haug A, Larsen B (1973) Polymannuronic acid 5-epimerase from the marine alga Pelvetia canaliculata. Acta Chem Scand 27:3592–3594 Morea A, Mathee K, Franklin MJ, Giacomini A, O’Regan M, Ohman DE (2001) Characterization of algG encoding C5-epimerase in the alginate biosynthetic gene cluster of Pseudomonas fluorescens. Gene 278:107–114 Mørch YA, Donati I, Strand BL, Skjåk-Bræk G (2007) Molecular engineering as an approach to design new functional properties of alginate. Biomacromolecules 8:2809–2814 Mørch YA, Holtan S, Donati I, Strand BL, Skjåk-Bræk G (2008) Mechanical properties of C-5 epimerized alginates. Biomacromolecules 9:2360–2368 Mrsny RJ, Lazazzera BA, Daugherty AL, Schiller NL, Patapoff TW (1994) Addition of a bacterial alginate lyase to purulent CF sputum in vitro can result in the disruption of alginate and modification of sputum viscoelasticity. Pulm Pharmacol 7:357–366
114
H. Ertesvåg et al.
Nyvall P, Corre E, Boisset C, Barbeyron T, Rousvoal S, Scornet D, Kloareg B, Boyen C (2003) Characterization of mannuronan C-5-epimerase genes from the brown alga Laminaria digitata. Plant Physiol 133:726–735 Ofstad R, Larsen B (1981) The effect on calcium-ion concentration on poly-D-mannuronate C-5epimerase. In: Levring TXth international seaweed symposium. Gothenburg, de Gruyter, Berlin, pp 485–493 Okazaki M, Furuya K, Tsukayama K, Nisizawa K (1982) Isolation and identification of alginic acid from a calcareous red alga Serraticarida maxima. Bot Mar 25:123–131 Osawa T, Matsubara Y, Muramatsu T, Kimura M, Kakuta Y (2005) Crystal structure of the alginate (Poly a-L-guluronate) lyase from Corynebacterium sp. at 1.2 Å resolution. J Mol Biol 345:1111–1118 Ramsey DM, Wozniak DJ (2005) Understanding the control of Pseudomonas aeruginosa alginate synthesis and the prospects for management of chronic infections in cystic fibrosis. Mol Microbiol 56:309–322 Ramstad MV, Ellingsen TE, Josefsen KD, Høidal HK, Valla S, Skjåk-Bræk G, Levine D (1999) Properties and action pattern of the recombinant mannuronan C-5-epimerase AlgE2. Enzyme Microb Technol 24:636–646 Rehm BH (1996) The Azotobacter vinelandii gene algJ encodes an outer-membrane protein presumably involved in export of alginate. Microbiology 142:873–880 Rehm BH, Boheim G, Tommassen J, Winkler UK (1994) Overexpression of algE in Escherichia coli: subcellular localization, purification, and ion channel properties. J Bacteriol 176:5639–5647 Rehm BH, Ertesvåg H, Valla S (1996) A new Azotobacter vinelandii mannuronan C-5-epimerase gene (algG) is part of an alg gene cluster physically organized in a manner similar to that in Pseudomonas aeruginosa. J Bacteriol 178:5884–5889 Robles-Price A, Wong TY, Sletta H, Valla S, Schiller NL (2004) AlgX is a periplasmic protein required for alginate biosynthesis in Pseudomonas aeruginosa. J Bacteriol 186:7369–7377 Rozeboom HJ, Bjerkan TM, Kalk KH, Ertesvåg H, Holtan S, Aachmann FL, Valla S, Dijkstra BW (2008) Structural and mutational characterization of the catalytic A-module of the mannuronan C-5-epimerase AlgE4 from Azotobacter vinelandii. J Biol Chem 235:23819–23828 Sadoff HL (1975) Encystment and germination in Azotobacter vinelandii. Bacteriol Rev 39:516–539 Shinabarger D (1993) Nucleotide sequence and expression of the Pseudomonas aeruginosa algF gene controlling acetylation of alginate. Mol Microbiol 9:1027–1035 Skjåk-Bræk G, Larsen B (1982a) A new assay for mannuronan C-5-epimerase activity. Carbohydr Res 103, 133–136 Skjåk-Bræk G, Larsen B (1982b) Purification of mannuronan C-5-epimerase by affinity chromatography on alginate-Sepharose. Carbohydr Res 103:137–140 Skjåk-Bræk G, Larsen B, Grasdalen H (1985) The role of O-acetyl groups in the biosynthesis of alginate by Azotobacter vinelandii. Carbohydr Res 145:169–174 Skjåk-Bræk G, Grasdalen H, Larsen B (1986a) Monomer sequence and acetylation pattern in some bacterial alginates. Carbohydr Res 154:239–250 Skjåk-Bræk G, Smidsrød O, Larsen B (1986b) Tailoring of alginates by enzymatic modifications in vitro. Int J Biol Macromol 8:330–336 Skjåk-Bræk G, Zanetti F, Paoletti S (1989) Effect of acetylation on some solution and gelling properties of alginate. Carbohydr Res 185:131–138 Sletmoen M, Skjåk-Bræk G, Stokke BT (2004) Indications of a processive mode of action of a non-degrading polysaccharide modifying enzyme obtained by single-molecular pair unbinding studies. Biomacromolecules 5:1288–1295 Smidsrød O, Draget KI (1996) Chemistry and physical properties of alginates. Carbohydr Eur 14:6–13 Stanford ECC (1883) On algin: a new substance obtained from some of the commoner species of marine algae. Chem News 96:254–257 Steigedal M, Sletta H, Moreno S, Mærk M, Christensen BE, Bjerkan T, Ellingsen TE, Espin G, Ertesvåg H, Valla S (2008) The Azotobacter vinelandii AlgE mannuronan C-5-epimerase
Enzymatic Alginate Modification
115
family is essential for the in vivo control of alginate monomer composition and for functional cyst formation. Environ Microbiol 10:1760–1770 Svanem BI, Skjåk-Bræk G, Ertesvåg H, Valla S (1999) Cloning and expression of three new Azotobacter vinelandii genes closely related to a previously described gene family encoding mannuronan C-5-epimerases. J Bacteriol 181:68–77 Svanem BI, Strand WI, Ertesvåg H, Skjåk-Bræk G, Hartmann M, Barbeyron T, Valla S (2001) The catalytic activities of the bifunctional Azotobacter vinelandii mannuronan C-5-epimerase and alginate lyase AlgE7 probably originate from the same active site in the enzyme. J Biol Chem 276:31542–31550 Tielen P, Strathmann M, Jaeger KE, Flemming HC, Wingender J (2005) Alginate acetylation influences initial surface colonization by mucoid Pseudomonas aeruginosa. Microbiol Res 160:165–176 Ullrich MS, Schergaut M, Boch J, Ullrich B (2000) Temperature-responsive genetic loci in the plant pathogen Pseudomonas syringae pv. glycinea. Microbiology 146:2457–2468 Vazquez A, Moreno S, Guzman J, Alvarado A, Espín G (1999) Transcriptional organization of the Azotobacter vinelandii algGXLVIFA genes: characterization of algF mutants. Gene 232:217–222 Windhues T, Borchard W (2003) Effect of acetylation on physio-chemical properties of bacterial and algal alginates in physiological sodium chloride solutions investigated with light scattering techniques. Carbohydr Polym 52:47–52 Yamasaki M, Moriwaki S, Miyake O, Hashimoto W, Murata K, Mikami B (2004) Structure and function of a hypothetical Pseudomonas aeruginosa protein PA1167 classified into family PL-7: a novel alginate lyase with a beta-sandwich fold. J Biol Chem 279:31863–31872
Alginate Gene Regulation Dennis E. Ohman
Contents 1 Introduction........................................................................................................................ 2 Hierarchical Regulation of the Alginate Operon ............................................................... 3 AlgR, a Two-Component Regulator .................................................................................. 4 AlgB, an NtrC-Family Two-Component Regulator ........................................................... 5 AmrZ, an Arc-Like DNA-Binding Protein ........................................................................ 6 Ancillary Regulators of the algD Operon .......................................................................... 7 Alternative Sigma Factor, s22............................................................................................. 8 Role of s22-MucAB and Regulated Proteolysis in a Cell Wall Stress Response ............... 9 Conclusions........................................................................................................................ References ................................................................................................................................
118 118 119 121 122 123 123 125 128 128
Abstract Alginate is an important virulence factor of Pseudomonas aeruginosa, and so our understanding of alginate gene regulation is best understood in this species. Expression of the algD operon for alginate biosynthesis is only highly expressed in mucoid clinical isolates that usually have pathoadaptive mucA mutations. The three major regulators of the algD promoter (PalgD) are AlgR, AlgB, and AmrZ. Each binds to DNA sites relatively far upstream from the start of algD transcription. AlgR and AlgB are two-component regulators. AmrZ is an Arc-like positive regulator of PalgD, but can also be a negative regulator. A global role for these regulators is also emerging. Expression of PalgD and the regulators are under the control of s22, an extracytoplasmic function alternative sigma factor. s22 activity is under posttranscriptional control by membrane-bound MucAB. Release of s22 sequestration can occur as a result of cell wall stress, which activates proteases, including AlgW (DegS-like) protease, to degrade MucA.
D.E. Ohman Department of Microbiology and Immunology, Virginia Commonwealth University Medical Center, P.O. Box 980678, Richmond, VA, 23298, USA McGuire Veterans Affairs Medical Center, Richmond, VA, 23298, USA e-mail: [email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_5, © Springer-Verlag Berlin Heidelberg 2009
117
118
1
D.E. Ohman
Introduction
Pseudomonas aeruginosa is a common opportunistic pathogen, and persons with the autosomal recessive disease cystic fibrosis (CF) are especially susceptible. A striking feature of strains infecting the CF pulmonary tract is the frequent conversion in vivo to a highly mucoid colony morphology, which is due to alginate overproduction (Linker and Jones 1966; Evans and Linker 1973). Alginate is a capsule-like exopolysaccharide that loosely adheres to the cells, and so most of it is found in the culture supernatant. Mucoid P. aeruginosa tends to persist in the lungs of CF patients, causing chronic bronchopulmonary infection (Govan and Harris 1986), and cannot be eliminated from the lungs despite aggressive antibiotic therapy (Frederiksen et al. 1997). The appearance of mucoid strains correlates with the formation of a bacterial biofilm containing microcolonies, the development of anti P. aeruginosa antibodies, inflammation, and a generally poor prognosis for the patient (Hoiby and Olling 1977; Lam et al. 1980; Pedersen 1992). The CF patient is eventually overwhelmed by the chronic infection as the number of P. aeruginosa in the sputum becomes as high as 108 cfu/ml (Seale et al. 1979). In addition, adults with chronic obstructive pulmonary disease have also been observed to have mucoid P. aeruginosa in their sputum samples (Brauer et al. 2008). There is evidence that the mucoid phenotype confers several selective advantages to the bacterial invader, which include increased resistance to phagocytosis (Schwarzmann and Boring III 1971; Baltimore and Mitchell 1982; Pedersen et al. 1992; Pier et al. 2001). Although alginate is synthesized by many species of Pseudomonas and by Azotobacter, the regulation of alginate in P. aeruginosa has been most intensively studied owing to the role of this capsule-like exopolysaccharide in pathogenesis. Thus, this review of alginate gene regulation will emphasize that which is known in P. aeruginosa.
2
Hierarchical Regulation of the Alginate Operon
All but one of the genes (i.e., algC) for alginate biosynthesis in P. aeruginosa are in a large operon (Chitnis and Ohman 1993), which is often called the “algD operon.” This 12-gene operon begins with algD and is located at 3.96-Mb on the 6.26-Mb physical genomic map of strain PAO1 (Stover et al. 2000). In typical P. aeruginosa strains, the algD operon is nearly silent, but it shows high activity in clinical mucoid strains (Deretic et al. 1987). The algD promoter is large and complex in that four major transcription factors are required for its expression in mucoid P. aeruginosa. There is a hierarchy of regulation in that an alternative sigma factor, s22 encoded by algT/algU, controls the algD promoter (PalgD) directly and also controls the expression of the other major regulators required for expression of the algD operon, i.e., AlgR, AlgB, and AlgZ (Wozniak and Ohman 1994). Each of these regulators binds to PalgD and is described in the following sections. Conversion to the mucoid phenotype and alginate overproduction as observed in clinical isolates is typically
Alginate Gene Regulation
119
due to mutations in mucA, which encodes a posttranscriptional negative regulator of s22 (Martin et al. 1993b; Boucher et al. 1997; Mathee et al. 1997).
3
AlgR, a Two-Component Regulator
The algR gene (PA5261), located at 5.92-Mb on the PAO1 chromosome map, encodes a 27.6-kDa protein of the two-component regulator family. AlgR (also called “AlgR1”) footprints to two segments far upstream of the algD promoter (–479 to –457 and –400 to –380 upstream of the transcription start site) and to a third binding site (–50 to –30) near the algD messenger RNA start site that is in an inverted orientation (Fig. 1); all binding sites share a consensus core sequence (ACCGTTCGTC) (Kato and Chakrabarty 1991; Mohr et al. 1991, 1992). AlgR contains a conserved N-terminal domain typical of response regulators (Fig. 2a) of two-component signal transduction systems (Deretic et al. 1989). The C-terminal domain has a LytTR domain (Fig. 2a) similar to LytR in Staphylococcus aureus (Nikolskaya and Galperin 2002), which is found in a variety of bacterial transcriptional regulators that bind to a specific DNA sequence pattern. Also under AlgR control is algC, which is not located in the algD operon. The algC gene encodes a phosphomannose/glucose mutase that is required for both alginate and lipopolysaccharide biosynthesis. AlgR binds to three regions upstream and downstream of the algC transcription start site (Fujiwara et al. 1993). Upstream of algR is algZ (PA5262, also called “fimS”), which encodes the 40-kDa sensory component that interacts with AlgR (Whitchurch et al. 1996; Yu et al. 1997). AlgZ has a LytS histidine kinase signal transduction domain related to that of LytS, a sensor in S. aureus (Fig. 2a). However, AlgZ lacks other recognizable nucleotide binding motifs that are typical of histidine protein kinases (Yu et al. 1997). The algZ–algR genes appear to form an operon, and it is under s22 positive control (Wozniak and Ohman 1994; Whitchurch et al. 1996). The start of algZ transcription is −73 bp from the start of translation (Wozniak and Ohman 1994). However, AlgZ is not required for alginate production, suggesting that AlgR has σ22 AlgR 0) 50 -3
+367
–(
+1
–(
00 -2 6 27 4) 422 4)
AmrZ AlgB
–( 3
80 )
AlgR
00 -3
–( 4
–(
47 9-
45 7)
AlgR
algD-A
Fig. 1 Map of the algD promoter (PalgD), which controls expression of the 12-gene operon (algD, 8, 44, K, E, G, X, L, I, J, F, A) for alginate biosynthesis in Pseudomonas aeruginosa. The closed arrowhead at +1 base pairs indicates the start of transcription, which is under s22 control. The relative locations of the upstream binding sites for major regulators AlgR, AlgZ, and AlgB are indicated. The start of algD translation is at +367 base pairs from the start of transcription
120
D.E. Ohman
Fig. 2 Gene maps of the major PalgD regulators: a AlgZ-AlgR, b AlgB-KinB, and c AmrZ. All are under the transcriptional control of s22. The names and locations of functional domains, obtained from the Kyoto Encyclopedia of Genes and Genomes (http://www.genome.jp/), are indicated on each open reading frame and are described in the text
both phosphorylation-dependent and phosphorylation-independent mechanisms of gene activation (Whitchurch et al. 1996, 2002; Yu et al. 1997; Ma et al. 1998). Even when the predicted phosphorylated residue of AlgR (Asp-54) was mutated to asparagine, alginate production was still activated (Ma et al. 1998). Although AlgZ is not required for alginate production in mucoid (mucA−) P. aeruginosa, AlgZ and AlgR are both required for type IV pilin mediated motility (i.e., twitching motility) on a solid surface (Whitchurch et al. 1996, 2002). The phosphorylation of AlgR at Asp-54 is essential for direct activation of the fimU-pilVWXY1Y2E operon, which is required for the assembly and export of a functional type IV pilus (Belete et al. 2008). AlgR also controls hydrogen cyanide production (Carterson et al. 2004) and activates expression of approximately 58 genes as shown by microarray analysis (Lizewski et al. 2004). In addition, AlgR represses the Rhl quorum-sensing system in nonmucoid biofilms but not in planktonic culture, suggesting a contact-dependent regulatory mechanism (Morici et al. 2007). In all, this suggests a more global role for AlgR in P. aeruginosa than strictly alginate gene regulation.
Alginate Gene Regulation
4
121
AlgB, an NtrC-Family Two-Component Regulator
A mutation in the algB gene blocks expression of PalgD in mucoid (i.e., mucA−) P. aeruginosa (Goldberg and Ohman 1984, 1987). The algB gene (PA5483) maps at 6.17-Mb on the chromosome and like algR encodes a two-component regulator (Wozniak and Ohman 1991; Goldberg and Dahnke 1992). Early attempts to demonstrate a direct interaction between AlgB and PalgD were unsuccessful, suggesting that AlgB may control PalgD indirectly (Woolwine and Wozniak 1999). However, recent studies have shown that AlgB binds within a region located between –274 and –224 (Fig. 1) relative to the start of PalgD transcription (Leech et al. 2008). Thus, AlgB belongs to a subclass of NtrC family proteins that can activate promoters that utilize a sigma factor other than s54 in that the PalgD utilizes s22 in mucoid P. aeruginosa. A footprint analysis has not yet been reported to determine the binding site of AlgB. The 49-kDa AlgB has an N-terminal response regulator domain (Fig. 2b) with the conserved Asp-16, Asp-59, and Lys-109 residues that are typically involved in phosphorylation. The central portion of AlgB has a s54 interaction domain and two Walker boxes (probably representing an AAA+ ATPase domain), which puts AlgB in the NtrC class of transcription enhancer-binding proteins (Wozniak and Ohman 1991). Although NtrC class regulators typically utilize s54 (sN) promoters, a mutation in rpoN (encoding s54) in mucoid (i.e., mucA-defective) P. aeruginosa has no effect on algD transcription or the mucoid phenotype (Mohr et al. 1990; Totten et al. 1990). The C-terminus of AlgB contains a helix-turn-helix for DNA binding that belongs to the Fis family of regulatory proteins (Fig. 2b). Immediately downstream of algB is kinB (PA5484), which encodes a 66-kDa histidine protein kinase that is the cognate environmental sensor of AlgB (Ma et al. 1997). The algB–kinB genes form an operon that is under s22 control, although the algB transcriptional start site has not yet been determined. The N-terminus of AlgB has two transmembrane domains that form a relatively large periplasmically localized loop as shown by PhoA fusion analysis (Ma et al. 1997). This loop is often the “sensory” domain of histidine protein kinases, and AlgB contains a cytochrome b-562 domain within it (Fig. 2b). In the cytochrome, this domain binds a single heme prosthetic group. Centrally, there is a HAMP domain, which is a cytoplasmic helical linker domain common to many receptor histidine kinase and methyl-accepting signaling proteins. Next is a PAS fold, which has been shown in other proteins to act as a sensor for oxygen and redox. There is a histidine kinase A domain, which is similar to that found in the signal-transducing histidine kinase, CheA. This domain contains the conserved “H box” where autophosphorylation at His-385 is predicted to occur. The C-terminus contains an HATPase c domain that is structurally related to ATPase domains of histidine kinases, and this contains the conserved N, D/F, and G boxes. To show experimentally that KinB has histidine protein kinase activity, a purified KinB derivative containing the C-terminal cytoplasmic region was demonstrated in vitro to undergo progressive autophosphoryation in the presence of [g-32P]ATP, and it was shown that the phosphoryl label could be rapidly transferred
122
D.E. Ohman
to purified AlgB (Ma et al. 1997). This reaction is blocked by substitutions of the residues conserved in histidine protein kinases. This provides evidence that autophosphorylation occurs at His-385 with transfer of the phosphate to Asp-59 in the response regulatory domain of AlgB (Ma et al. 1997). The presence of this active two-component regulatory system suggests that AlgB regulates alginate gene expression in response to environmental or physiological changes. However, KinB is not required for the mucoid phenotype in a mucA-defective background (Ma et al. 1998). Even algB alleles expressing proteins blocked for phosphorylation, with a D59N substitution or even with the N-terminal response regulatory domain (i.e., 145 residues) deleted, could still restore the mucoid phenotype to an algB null mutant of P. aeruginosa. Nevertheless, there is evidence that KinB plays a regulatory role in the cell by phosphorylating AlgB. A microarray analysis has shown that as many as 89 genes require both AlgB and KinB for maximal expression, and thus AlgB has a global role beyond alginate gene regulation (Leech et al. 2008).
5
AmrZ, an Arc-Like DNA-Binding Protein
The amrZ gene (PA3385, located at 3.79-Mb on the PAO1 map) was originally called “algZ” as an alginate regulator (Baynham and Wozniak 1996; Baynham et al. 1999), but was later renamed “amrZ” for “alginate and motility regulator” (Baynham et al. 2006). When amrZ is mutated in a mucoid strain, there is no alginate production or detectable algD transcription (Baynham et al. 1999). A footprint analysis shows that AmrZ binds to a 36-bp sequence that contains an 8-bp direct repeat (GCCATTAC) and is centered at −282 bp relative to the algD start of transcription (Baynham and Wozniak 1996) (Fig. 1). The amrZ gene is under s22 positive control and encodes a 12-kDa protein that is a DNA-binding protein of the ribbon–helix–helix family showing homology to the Mnt and Arc repressors of Salmonella bacteriophage 22 (Baynham and Wozniak 1996; Baynham et al. 1999). These proteins have an N-terminal b-sheet involved in recognizing and binding DNA. AmrZ contains an Arc-like DNA binding domain (Fig. 2c), and mutations in AmrZ within this b-sheet result in loss of DNA binding (Ramsey et al. 2005; Baynham et al. 2006). AmrZ can be either a gene activator or a gene repressor. Like AlgR, AmrZ has a role in type IV pilus biosynthesis and thus twitching motility. A mutation in amrZ results in no type IV pilin expression, and the amount of PilA (the major pilin subunit) on the surface of the cells is much reduced (Baynham et al. 2006). In contrast, AmrZ binds two sites at its own promoter, where it functions as an autorepressor (Ramsey et al. 2005). AmrZ is also a negative regulator of flagellum biosynthesis. In mucoid P. aeruginosa, where amrZ shows high expression (due to deregulated s22 activity), AmrZ blocks flagellum biosynthesis by binding to the promoter of fleQ, which encodes the master regulator of the flagella regulon (Tart et al. 2006). Thus, AmrZ is a multifunctional regulator of alginate and motility.
Alginate Gene Regulation
6
123
Ancillary Regulators of the algD Operon
Other regulators of algD expression have also been described. AlgQ (AlgR2) is a positive regulator of nucleoside diphosphokinase, which is necessary for the formation of GDP-mannose, a precursor in alginate biosynthesis (Kim et al. 1998). AlgP (AlgR3) is a highly basic histone-like protein required for normal algD expression (Deretic and Konyecsni 1989, 1990; Kato et al. 1990), but it does not appear to bind PalgD. Integration host factor is a global regulator that has been shown to play a role in algD expression (Wozniak and Ohman 1993; Delic-Attree et al. 1996, 1997). RpoS, the stationary-phase sigma factor, is also required for high-level alginate production (Suh et al. 1999). CysB, required for cysteine biosynthesis, can act as an activator of algD expression (Delic-Attree et al. 1997). Unlike mucA mutants, a mucoid strain with an undefined muc-23 mutation depends on s54 (RpoN) for expression of PalgD instead of s22 (AlgT/AlgU)(Boucher et al. 2000). The intracellular level of the bacterial second messenger bis-(3¢–5¢)-cyclic-GMP (c-di-GMP) also regulates alginate production. Alg44, encoded by the algD operon, has a PilZ domain that functions as a c-di-GMP receptor, and substitutions in the PilZ conserved residues results in loss of c-di-GMP binding and the ability to produce alginate in P. aeruginosa (Merighi et al. 2007). Alg44 is an inner-membrane protein required for alginate polymerization, and its PilZ domain is located in a cytoplasmically localized domain (Oglesby et al. 2008). Growth conditions may affect alginate production. The environment of the CF lung has a high osmolarity, and dehydration has been reported to contribute to the activation of the algD promoter (Berry et al. 1989; DeVault et al. 1991). The addition of calcium (1.0 and 10 mM CaCl2) to the culture results in tenfold thicker biofilms and eightfold increase in the expression of algD in the mucoid CF strain, FRD1 (Sarkisova et al. 2005), but the regulatory mechanisms are not yet understood. Anaerobic growth conditions have also been shown to stimulate alginate production in normally nonmucoid P. aeruginosa (Hassett 1996).
7
Alternative Sigma Factor, s22
The algT gene (PA0762, also called “algU”) is located at 0.83 Mb on the PAO1 genome map and encodes a 22-kDa alternative sigma factor called “s22” (also called “AlgT,” “AlgU,” “sE”) (Deretic et al. 1994; DeVries and Ohman 1994). Early studies showed that algT is essential for alginate production in mucoid strains and that it is genetically linked to the “switch to mucoid” gene (originally called “algS” and now called “mucA”) (Flynn and Ohman 1988a, b). Overexpressed from a multicopy plasmid, algT can activate the mucoid phenotype in typical nonmucoid strains of P. aeruginosa like PAO1 and in several other Pseudomonas species (Goldberg et al. 1993). s22 shows sigma factor activity in in vitro transcription assays (Hershberger et al. 1995). Escherichia coli sE is sufficiently similar to P. aeruginosa s22 that the E. coli rpoE gene can complement an algT(algU) mutation in P. aeruginosa, restoring the mucoid phenotype (Yu et al. 1995). s22 promotes transcriptional activity at promoters with the
124
D.E. Ohman
consensus sequence GAACTT (−35) … TCtga (−10) (DeVries and Ohman 1994; Govan and Deretic 1996). PalgD in P. aeruginosa matches this consensus. The algT product positively regulates its own transcription at two starts of transcription that also have the s22 consensus sequence (DeVries and Ohman 1994; Schurr et al. 1995) The algT gene is in an operon with four other downstream genes called “mucA– mucB–mucC–mucD” (Fig. 3), which are named “muc“ for the mucoid phenotype caused by mutations in this region (Martin et al. 1993b). MucA (21 kDa) is an inner-membrane protein with one transmembrane domain, and it functions to posttranslationally control s22 activity (Mathee et al. 1997). MucA acts as an anti-sigma factor that binds s22 and can inhibit its ability to transcribe from PalgD in in vitro transcription assays (Schurr et al. 1996; Xie et al. 1996). As mentioned already, mucoid conversion observed in CF clinical isolates of P. aeruginosa is typically due to mutations in mucA that occur in vivo (Martin et al. 1993b; Boucher et al. 1997). A common mutation, called the “mucA22 allele,” has a single base pair deletion in a string of five Gs that results in a prematurely truncated MucA. The nature of the selection pressure for mucoid conversion in the CF lung has received much speculation. Treatment of a biofilm culture of nonmucoid P. aeruginosa PAO1 with activated polymorphonuclear leukocytes can result in mucoid conversion (Mathee et al. 1999). When CF isolates are cultured in the laboratory, the mucoid phenotype is typically unstable, especially under nonaerated conditions (Ohman and Chakrabarty 1981). This is usually due to mutations in the algT gene instead of by reversion to restore the wild-type mucA allele (DeVries and Ohman 1994; Schurr et al. 1994). MucB (PA0764, also called “AlgN”) is a 34.6-kDa negative regulator of s22 activity (Goldberg et al. 1993; Martin et al. 1993a). MucB is a periplasmic protein (Mathee et al. 1997) that binds the periplasmic domain of MucA (L.F. Wood and D.E. Ohman, unpublished data). Mutation in mucB leads to a partially mucoid phenotype, probably owing to the lack of MucA-MucB interactions (Martin et al. 1993a). Thus, MucA–MucB apparently forms a signal transduction complex that normally keeps s22 transcriptional activity low until it is affected by an appropriate environmental stress signal. However, in the CF pulmonary environment, there is a strong selection for mucA defects to bypass this control circuit and deregulate the negative control over s22 transcriptional activity. The 16-kDa MucC open reading frame appears to be translationally coupled to MucB but its role is unclear (Fig. 3). MucD (PA0766) has high similarity to E. coli HtrA (DegP) (Fig. 3), a periplasmic serine protease involved in the proteolysis of abnormal proteins and that is required algT σ22
mucA
mucB
mucC
mucD 1 kb
Fig. 3 Map of the algT(algU) operon of P. aeruginosa. The two closed arrowheads upstream indicate starts of transcription under autoregulatory control by the algT product, s22. MucA-MucB posttranscriptionally control the activity of s22. The closed arrowhead within mucC represents a strong secondary promoter (not under s22 control) for expression of mucD, which encodes a periplasmic HtrA-like protease
Alginate Gene Regulation
125
for resistance to oxidative and heat stress (Ohman et al. 1996; Boucher et al. 1997). Mutation in mucD results in a semimucoid phenotype, especially on Pseudomonas isolation agar (Boucher et al. 1997; Wood and Ohman 2006). The mucD gene has a secondary promoter, independent of s22, that is curiously located in the middle of mucC (Wood and Ohman 2006). Thus, the phenotype of a mucC mutation is complicated by its effects on the mucD promoter and mucD expression. A mutant expressing a single-copy mucD allele encoding MucD altered in its serine protease motif (S217A) is defective in temperature resistance and alginate gene regulation, and so it is a lack of MucD proteolytic activity that activates the mucoid phenotype (Wood and Ohman 2006).
8
Role of s22-MucAB and Regulated Proteolysis in a Cell Wall Stress Response
P. aeruginosa s22 belongs to the extracytoplasmic function subfamily of sigma factors, which are distantly related to s70 and generally play roles in stress responses (Lonetto et al. 1994; Missiakas and Raina 1998; Helmann 2002). E. coli sE, a well-characterized extracytoplasmic function sigma factor, is sequestered by RseAB proteins (which resemble MucAB) and together they form an envelope-stress response system (Alba and Gross 2004). The sE–RseAB complex sequesters the sigma factor until envelope stress (e.g., misfolded envelope proteins) signals DegS protease and YaeL/RseP protease to sequentially degrade the anti-sigma factor RseA in a process called “regulated intramembrane proteolysis,” which derepresses sigma factor activity (Alba and Gross 2004). Given this striking similarity to the sE-RseAB complex, a broader role for s22 in P. aeruginosa is emerging as an extracytoplasmic cell wall stress response system. P. aeruginosa s22 controls several genes other than those for alginate production, including rpoH, a gene encoding s32, a homologue of the major heat-shock sigma factor (Schurr and Deretic 1997). A proteomic study, using a 2D-gel analysis of all cellular proteins, showed that unregulated s22 in a mucoid PAO1 mucA22 strain causes elevated expression of dsbA, encoding disulfide bond isomerase (Malhotra et al. 2000). A genomic analysis using DNA microarrays showed that the expression of numerous genes having a wide array of functions is coinduced with alginate genes in a mucoid PAO1 mucA strain compared with the wild type (Firoved and Deretic 2003). Thus, like sE–RseAB in E. coli, the P. aeruginosa s22–MucAB membrane complex could be a signal transduction mechanism that responds to cell stress to activate the expression of many genes. To identify stress stimuli that cause transcriptional induction of PalgD, a plate bioassay has been employed where a plasmid-borne algD promoter is fused to a chloramphenicol acetyltransferase cartridge (PalgD-cat) to provide chloramphenicol resistance as a marker for PalgD activation (Fig. 4a). This revealed that antibiotics affecting peptidoglycan synthesis and other agents disruptive to the cell wall have a strong effect on PalgD induction in wild-type P. aeruginosa PAO1 (Wood et al. 2006). For instance, treatment of wild-type P. aeruginosa PAO1 carrying the PalgD-cat plasmid leads to a ring of growth around d-cycloserine, a peptidoglycan synthesis
126
D.E. Ohman
Fig. 4 Bioassay for environmental factors that activate PalgD expression. a The broad host range plasmid pKK61 contains the complete PalgD region fused to a cat (chloramphenicol acetyltransferase) reporter gene, which confers chloramphenicol resistance when the plasmid-encoded PalgD is activated. b Agar plates containing chloramphenicol are spread with P. aeruginosa strains carrying pKK61. Wild-type PAO1 (left) shows no growth except at the center where D-cycloserine was deposited, which activates PalgD owing to its induction of a cell wall stress response. An algW mutant of PAO1 (right), defective in expression of AlgW (DegS-like) protease, fails to show PalgD-cat expression in the presence of D-cycloserine, indicating a role for this protease in the cell wall stress response that derepresses s22 activity
inhibitor (Fig. 4b). This induction is blocked by mutations in the genes for s22, AlgB, and AlgR. A transcriptome analysis using microarrays verified that sublethal levels of d-cycloserine activate expression of genes in the algD operon. d-Cycloserine also activated many of the same genes, seen in earlier microarray assays (Firoved and Deretic 2003), that are associated with s22 derepression (Wood et al. 2006). Thus, in wild-type nonmucoid P. aeruginosa, the s22 regulon responds to cell wall stress. There is also evidence for regulated intramembrane proteolysis (RIP) control of s22 activity as seen with sE in E. coli. A gene called “algW” (PA4446 at 4.9 Mb on the PAO1 map), originally described for its mutant phenotype affecting alginate production (Boucher et al. 1996), is the functional homologue of DegS protease in P. aeruginosa. In the bioassay described above, loss of AlgW protease blocks PalgD induction by d-cycloserine (Fig. 4b), and overexpression of algW in PAO1 results in a mucoid phenotype even in the absence of cell wall stress (Wood et al.
Alginate Gene Regulation
127
2006). These data suggest that AlgW protease plays a major role in s22 activation. Other evidence for RIP in alginate gene regulation came about through a chance discovery following random transposon mutagenesis that overexpression of PA4033 (called “mucE”) causes mucoid conversion in P. aeruginosa (Qiu et al. 2007). PA4033 is a 9.5-kDa envelope lipoprotein, and its overproduction and accumulation derepresses s22 in a manner that may be similar to that seen in E. coli where overexpression of porins causes sE derepression (Alba and Gross 2004). The effect of MucE overexpression on alginate production also requires the AlgW protease and another protease (termed “MucP”), which is a homologue of RseP/YaeL protease in E. coli; evidence for MucA degradation is also observed (Qiu et al. 2007). A model for PalgD activation is shown in Fig. 5. MucAB sequesters s22 in the inner membrane, thus preventing its interaction with promoters and RNA polymerase, which keeps PalgD nearly silent. In the CF pulmonary environment there is a selection for mucA mutations that make MucA defective, thus releasing s22 to cause deregulation of the system, and allowing maximal expression of PalgD. In wild-type (MucA+) P. aeruginosa, cell wall stress (e.g., peptidoglycan inhibitors, misfolded porins) activate a RIP cascade in which the AlgW and YaeL/MucP proteases degrade MucA to release s22, thus allowing PalgD expression (Fig. 5).
OM Cell wall Stress
MucD MucB
MucB
in vivo mutation AlgW
MucB
‘MucA
IM
MucA’
MucA
σ22
YaeL
MucA’
σ22
σ22 algD-A PalgD Active
algD-A PalgD Silent
algD-A PalgD Active
Fig. 5 Model for PalgD activation in P. aeruginosa. In the center is shown MucAB sequestering s22 at the inner membrane to prevent interaction with promoters and RNA polymerase, which keeps PalgD nearly silent. On the left is shown the effect of a mucA mutation that truncates MucA prematurely (a common event in the cystic fibrosis pulmonary environment), and thus release of s22 causing maximal expression of PalgD. On the right is shown the effect of cell wall stress (e.g., peptidoglycan inhibitors, misfolded porins) on wild-type P. aeruginosa, which activates the AlgW and YaeL/MucP proteases to degrade MucA and thus release s22 for PalgD expression
128
D.E. Ohman
Other proteases are also being examined for effects on alginate gene regulation. A mutation in the Prc-like protease (PA3257) of P. aeruginosa, originally reported to affect alginate production (Reiling et al. 2005), causes poor induction of PalgD-cat by d-cycloserine, suggesting that it also can play a role in the response to cell wall stress (Wood et al. 2006). MucD, the periplasmic HtrA-like protease is encoded by the algT operon, and its function may be to degrade misfolded proteins in that compartment. When MucD is missing by mutation, misfolded proteins in the periplasm signal the RIP cascade, activation of the AlgW protease cascade, and expression of PalgD (Fig. 5). There is new evidence that three other proteases, ClpP (PA1801), ClpX (PA1802), and ClpP2 (PA3326), may have a role in the degradation of MucA fragments in the cytoplasm (Qiu et al. 2008).
9
Conclusions
Activation of the algD promoter and thus alginate biosynthesis is remarkably complex and is linked to a system for stress response. Two regulators (AlgR and AlgB) are proteins of the two-component regulator family, and yet their cognate sensors are not required for PalgD expression in mucoid P. aeruginosa. Two alginate regulators (AlgR and AmrZ) are also involved in motility. The fact that the PalgD region is about a kilobase in size and requires multiple transcriptional activators suggests the formation of a high-order loop structure allowing for multivalent contacts between the multiple transcription factors. The requirement for three transcriptional activators (AlgR, AlgB, AmrZ) to activate PalgD suggests that RNA polymerase may be contacted simultaneously by all three. It is unusual for a bacterial promoter to be controlled by the binding of three activators, especially when it is positioned so far upstream from the transcription start site. The binding of integration host factor, which has strong DNA-binding properties, may help to promote the interactions of RNA polymerase with the three regulatory proteins. In addition, it is intriguing that PalgD has such a long (367-nucleotide) 5¢ untranslated sequence (Fig. 1), but its role in regulating PalgD expression has as yet not been explored. Acknowledgements This work was supported by Veterans Administration Medical Research Funds, by a grant from the Cystic Fibrosis Foundation, and by Public Health Service grants AI-19146 and T32-AI-007617 from the National Institute of Allergy and Infectious Diseases.
References Alba BM, Gross CA (2004) Regulation of the Escherichia coli sigma-dependent envelope stress response. Mol Microbiol 52:613–619 Baltimore RS, Mitchell M (1982) Immunologic investigations of mucoid strains of Pseudomonas aeruginosa: comparison of susceptibility to opsonic antibody in mucoid and nonmucoid strains. J Infect Dis 141:238–247
Alginate Gene Regulation
129
Baynham P, Wozniak D (1996) Identification and characterization of AlgZ, an AlgT-dependent DNA-binding protein required for Pseudomonas aeruginosa algD transcription. Mol Microbiol 22:97–108 Baynham PJ, Brown AL, Hall LL, Wozniak DJ (1999) Pseudomonas aeruginosa AlgZ, a ribbonhelix-helix DNA-binding protein, is essential for alginate synthesis and algD transcriptional activation. Mol Microbiol 33:1069–1080 Baynham PJ, Ramsey DM, Gvozdyev BV, Cordonnier EM, Wozniak DJ (2006) The Pseudomonas aeruginosa ribbon-helix-helix DNA-binding protein AlgZ (AmrZ) controls twitching motility and biogenesis of type IV pili. J Bacteriol 188:132–140 Belete B, Lu H, Wozniak DJ (2008) Pseudomonas aeruginosa AlgR regulates type IV pilus biosynthesis by activating transcription of the fimU-pilVWXY1Y2E operon. J Bacteriol 190:2023–2030 Berry A, DeVault JD, Chakrabarty AM (1989) High osmolarity is a signal for enhanced algD transcription in mucoid and nonmucoid Pseudomonas aeruginosa strains. J. Bacteriol. 171:2312–2317 Boucher JC, Martinez-Salazar J, Schurr MJ, Mudd MH, Yu H, Deretic V (1996) Two distinct loci affecting conversion to mucoidy in Pseudomonas aeruginosa in cystic fibrosis encode homologs of the serine protease HtrA. J Bacteriol 178:511–523 Boucher JC, Yu H, Mudd MH, Deretic V (1997) Mucoid Pseudomonas aeruginosa in cystic fibrosis: characterization of muc mutations in clinical isolates and analysis of clearance in a mouse model of respiratory infection. Infect Immun 65:3838–3846 Boucher JC, Schurr MJ, Deretic V (2000) Dual regulation of mucoidy in Pseudomonas aeruginosa and sigma factor antagonism. Mol Microbiol 36:341–351 Brauer AL, Wozniak DJ, Murphy TF (2008) Characterization of mucoid Pseudomonas aeruginosa isolated in chronic obstructive pulmonary disease. In: Abstracts of the 108th general meeting of the American Society for Microbiology. American Society for Microbiology, Boston Carterson AJ, Morici LA, Jackson DW, Frisk A, Lizewski SE, Jupiter R, Simpson K, Kunz DA, Davis SH, Schurr JR, Hassett DJ, Schurr MJ (2004) The transcriptional regulator AlgR controls cyanide production in Pseudomonas aeruginosa. J Bacteriol 186:6837–6844 Chitnis CE, Ohman DE (1993) Genetic analysis of the alginate biosynthetic gene cluster of Pseudomonas aeruginosa shows evidence of an operonic structure. Mol Microbiol 8:583–590 Delic-Attree I, Toussaint B, Froger A, Willison JC, Vignais PM (1996) Isolation of an IHF-deficient mutant of a Pseudomonas aeruginosa mucoid isolate and evaluation of the role of IHF in algD gene expression. Microbiology 142:2785–2793 Delic-Attree I, Toussaint B, Garin J, Vignais PM (1997) Cloning, sequence and mutagenesis of the structural gene of Pseudomonas aeruginosa CysB, which can activate algD transcription. Mol Microbiol 24:1275–1284 Deretic V, Konyecsni WM (1989) Control of mucoidy in Pseudomonas aeruginosa: transcriptional regulation of algR and identification of the second regulatory gene, algQ. J Bacteriol 171:3680–3688 Deretic V, Konyecsni WM (1990) A procaryotic regulatory factor with a histone H1-like carboxyterminal domain: clonal variation of repeats within algP, a gene involved in regulation of mucoidy in Pseudomonas aeruginosa. J Bacteriol 172:5544–5554 Deretic V, Gill JF, Chakrabarty AM (1987) Gene algD coding for GDP-mannose dehydrogenase is transcriptionally activated in mucoid Pseudomonas aeruginosa. J Bacteriol 169:351–358 Deretic V, Dikshit R, Konyecsni M, Chakrabarty AM, Misra TK (1989) The algR gene, which regulates mucoidy in Pseudomonas aeruginosa, belongs to a class of environmentally responsive genes. J Bacteriol 171:1278–1283 Deretic V, Schurr MJ, Boucher JC, Deretic V, Schurr MJ, Boucher JC, Martin DW (1994) Conversion of Pseudomonas aeruginosa to mucoidy in cystic fibrosis: environmental stress and regulation of bacterial virulence by alternative sigma factors. J Bacteriol 176:2773–2780 DeVault JD, Hendrickson W, Kato J, Chakrabarty AM (1991) Environmentally regulated algD promoter is responsive to the cAMP receptor protein in Escherichia coli. Mol Microbiol 5:2503–2509
130
D.E. Ohman
DeVries CA, Ohman DE (1994) Mucoid to nonmucoid conversion in alginate-producing Pseudomonas aeruginosa often results from spontaneous mutations in algT, encoding a putative alternative sigma factor, and shows evidence for autoregulation. J Bacteriol 176:6677–6687 Evans LR, Linker A (1973) Production and characterization of the slime polysaccharide of Pseudomonas aeruginosa. J Bacteriol 116:915–924 Firoved AM, Deretic V (2003) Microarray analysis of global gene expression in mucoid Pseudomonas aeruginosa. J Bacteriol 185:1071–1081 Flynn JL, Ohman DE (1988a) Cloning of genes from mucoid Pseudomonas aeruginosa which control spontaneous conversion to the alginate production phenotype. J Bacteriol 170:1452–1460 Flynn JL, Ohman DE (1988b) Use of a gene replacement cosmid vector for cloning alginate conversion genes from mucoid and nonmucoid Pseudomonas aeruginosa strains: algS controls expression of algT. J Bacteriol 170:3228–3236 Frederiksen B, Koch C, Høiby N (1997) Antibiotic treatment of initial colonization with Pseudomonas aeruginosa postpones chronic infection and prevents deterioration of pulmonary function in cystic fibrosis. Pediatr Pulmonol 23:330–335 Fujiwara S, Zielinski NA, Chakrabarty AM (1993) Enhancer-like activity of AlgR1-binding site in alginate gene activation: positional, orientational, and sequence specificity. J Bacteriol 175:5452–5459 Goldberg JB, Dahnke T (1992) Pseudomonas aeruginosa AlgB, which modulates the expression of alginate, is a member of the NtrC subclass of prokaryotic regulators. Mol Microbiol 6:59–66 Goldberg JB, Ohman DE (1984) Cloning and expression in Pseudomonas aeruginosa of a gene involved in the production of alginate. J Bacteriol 158:1115–1121 Goldberg JB, Ohman DE (1987) Construction and characterization of Pseudomonas aeruginosa algB mutants: role of algB in high-level production of alginate. J Bacteriol 169:1593–1602 Goldberg JB, Gorman WL, Flynn JL, Ohman DE (1993) A mutation in algN permits trans-activation of alginate production by algT in Pseudomonas species. J Bacteriol 175:1303–1308 Govan JRW, Deretic V (1996) Microbial pathogenesis in cystic fibrosis: mucoid Pseudomonas aeruginosa and Burkholderia cepacia. Microbiol Rev 60:539–574 Govan JRW, Harris GS (1986) Pseudomonas aeruginosa and cystic fibrosis: unusual bacterial adaptation and pathogenesis. Microbiol Sci 3:302–308 Hassett DJ (1996) Anaerobic production of alginate by Pseudomonas aeruginosa: alginate restricts diffusion of oxygen. J Bacteriol 178:7322–7325 Helmann JD (2002) The extracytoplasmic function (ECF) sigma factors. Adv Microb Physiol 46:47–110 Hershberger CD, Ye RW, Parsek MR, Xie Z, Chakrabarty AM (1995) The algT (algU) gene of Pseudomonas aeruginosa, a key regulator involved in alginate biosynthesis, encodes an alternative s factor (sE). Proc Natl Acad Sci U S A 92:7941–7945 Hoiby N, Olling S (1977) Pseudomonas aeruginosa infection in cystic fibrosis: bactericidal effect of serum from normal individuals and patients with cystic fibrosis on P. aeruginosa strains from patients with cystic fibrosis or other diseases. Acta Pathol Microbiol Scand C 85:107–114 Kato J, Chakrabarty AM (1991) Purification of the regulatory protein AlgR1 and its binding in the far upstream region of the algD promoter in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 88:1760–1764 Kato J, Misra TK, Chakrabarty AM (1990) AlgR3, a protein resembling eukaryotic histone H1, regulates alginate synthesis in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 87:2887–2891 Kim H, Schlictman D, Shankar S, Xie Z, Chakrabarty A, Kornberg A (1998) Alginate, inorganic polyphosphate, GTP and ppGpp synthesis co-regulated in Pseudomonas aeruginosa: implications for stationary phase survival and synthesis of RNA/DNA precursors. Mol Microbiol 27:717–725 Lam J, Chan R, Lam K, Costerton JRW (1980) Production of mucoid microcolonies by Pseudomonas aeruginosa within infected lungs in cystic fibrosis. Infect Immun 28:546–556 Leech AJ, Sprinkle A, Wood L, Wozniak DJ, Ohman DE (2008) The NtrC family regulator AlgB, which controls alginate biosynthesis in mucoid Pseudomonas aeruginosa, binds directly to the algD promoter. J Bacteriol 190:581–589
Alginate Gene Regulation
131
Linker A, Jones RS (1966) A new polysaccharide resembling alginic acid isolated from Pseudomonads. J Biol Chem 241:3845–3851 Lizewski SE, Schurr JR, Jackson DW, Frisk A, Carterson AJ, Schurr MJ (2004) Identification of AlgR-regulated genes in Pseudomonas aeruginosa by use of microarray analysis. J Bacteriol 186:5672–5684 Lonetto MA, Brown KL, Rudd KE, Buttner MJ (1994) Analysis of Streptomyces coelicolor SigE gene reveals the existence of a subfamily of eubacterial RNA polymerase s factors involved in the regulation of extracytoplasmic functions. Proc Natl Acad Sci U S A 91:7573–7577 Ma S, Wozniak DJ, Ohman DE (1997) Identification of the histidine protein kinase KinB in Pseudomonas aeruginosa and its phosphorylation of the alginate regulator AlgB. J Biol Chem 272:17952–17960 Ma S, Selvaraj U, Ohman DE, Quarless R, Hassett DJ, Wozniak DJ (1998) Phosphorylationindependent activity of the response regulators AlgB and AlgR in promoting alginate biosynthesis in mucoid Pseudomonas aeruginosa. J Bacteriol 180:956–968 Malhotra S, Silo-Suh LA, Mathee K, Ohman DE (2000) Proteome analysis of the effect of mucoid conversion on global protein expression in Pseudomonas aeruginosa strain PAO1 shows induction of the disulfide bond isomerase, DsbA. J Bacteriol 182:6999–7006 Martin DW, Schurr MJ, Mudd MH, Deretic V (1993a) Differentiation of Pseudomonas aeruginosa into the alginate-producing form: inactivation of mucB causes conversion to mucoidy. Mol Microbiol 9:497–506 Martin DW, Schurr MJ, Mudd MH, Govan JRW, Holloway BW, Deretic V (1993b) Mechanism of conversion to mucoidy in Pseudomonas aeruginosa infecting cystic fibrosis patients. Proc Natl Acad Sci U S A 90:8377–8381 Mathee K, McPherson CJ, Ohman DE (1997) Posttranslational control of the algT (algU)-encoded s22 for expression of the alginate regulon in Pseudomonas aeruginosa and localization of its antagonist proteins MucA and MucB (AlgN). J Bacteriol 179:3711–3720 Mathee K, Ciofu O, Sternberg CK, Lindum P, Campbell J, Jensen P, Johnsen A, Givskov M, Ohman D, Molin S, Høiby N, Kharazmi A (1999) Mucoid conversion of Pseudomonas aeruginosa by hydrogen peroxide: a mechanism for virulence activation in the cystic fibrosis lung. Microbiology 145:1349–1357 Merighi M, Lee VT, Hyodo M, Hayakawa Y, Lory S (2007) The second messenger bis-(3¢–5¢)cyclic-GMP and its PilZ domain-containing receptor Alg44 are required for alginate biosynthesis in Pseudomonas aeruginosa. Mol Microbiol 65:876–895 Missiakas D, Raina S (1998) The extracytoplasmic function sigma factors: role and regulation. Mol Microbiol 28:1059–1066 Mohr CD, Martin DW, Konyecsni WM, Govan JRW, Lory S, Deretic V (1990) Role of the far-upstream sites of the algD promoter and the algR and rpoN genes in environmental modulation of mucoidy in Pseudomonas aeruginosa. J Bacteriol 172:6576–6580 Mohr CD, Hibler NS, Deretic V (1991) AlgR, a response regulator controlling mucoidy in Pseudomonas aeruginosa, binds to the FUS sites of the algD promoter located unusually far upstream from the mRNA start site. J Bacteriol 173:5136–5143 Mohr CD, Leveau JHJ, Krieg DP, Hibler NS, Deretic V (1992) AlgR-Binding sites within the algD promoter make up a set of inverted repeats separated by a large intervening segment of DNA. J Bacteriol 174:6624–6633 Morici LA, Carterson AJ, Wagner VE, Frisk A, Schurr JR, Honer zu Bentrup K, Hassett DJ, Iglewski BH, Sauer K, Schurr MJ (2007) Pseudomonas aeruginosa AlgR represses the Rhl quorum-sensing system in a biofilm-specific manner. J Bacteriol 189:7752–7764 Nikolskaya AN, Galperin MY (2002) A novel type of conserved DNA-binding domain in the transcriptional regulators of the AlgR/AgrA/LytR family. Nucleic Acids Res 30:2453–2459 Oglesby LL, Jain S, Ohman DE (2008) Membrane topology and roles of Pseudomonas aeruginosa Alg8 and Alg44 in alginate polymerization. Microbiology 154:1605–1615 Ohman DE, Chakrabarty AM (1981) Genetic mapping of chromosomal determinants for the production of the exopolysaccharide alginate in a Pseudomonas aeruginosa cystic fibrosis isolate. Infect Immun 33:142–148
132
D.E. Ohman
Ohman DE, Mathee K, McPherson CJ, DeVries CA, Ma S, Wozniak DJ, Franklin MJ (1996) Regulation of the alginate (algD) operon in Pseudomonas aeruginosa. In: Nakazawa T, Furukawa K, Haas D, Silver S (eds) Molecular biology of pseudomonads. American Society for Microbiology, Washington, pp 472–483 Pedersen SS (1992) Lung infection with alginate-producing, mucoid Pseudomonas aeruginosa in cystic fibrosis. APMIS Suppl 28:1–79 Pedersen SS, Moller H, Espersen F, Sørensen CH, Jensen T, Høiby N (1992) Mucosal immunity to Pseudomonas aeruginosa alginate in cystic fibrosis. Acta Pathol Microbiol Immunol Scand 100:326–334 Pier GB, Coleman F, Grout M, Franklin M, Ohman DE (2001) Role of alginate O acetylation in resistance of mucoid Pseudomonas aeruginosa to opsonic phagocytosis. Infect Immun 69:1895–1901 Qiu D, Eisinger VM, Rowen DW, Yu HD (2007) Regulated proteolysis controls mucoid conversion in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 104:8107–8112 Qiu D, Eisinger VM, Head NE, Pier GB, Yu HD (2008) ClpXP proteases positively regulate alginate overexpression and mucoid conversion in Pseudomonas aeruginosa. Microbiology 154:2119–2130 Ramsey DM, Baynham PJ, Wozniak DJ (2005) Binding of Pseudomonas aeruginosa AlgZ to sites upstream of the algZ promoter leads to repression of transcription. J Bacteriol 187:4430–4443 Reiling SA, Jansen JA, Henley BJ, Singh S, Chattin C, Chandler M, Rowen DW (2005) Prc protease promotes mucoidy in mucA mutants of Pseudomonas aeruginosa. Microbiology 151:2251–2261 Sarkisova S, Patrauchan MA, Berglund D, Nivens DE, Franklin MJ (2005) Calcium-induced virulence factors associated with the extracellular matrix of mucoid Pseudomonas aeruginosa biofilms. J Bacteriol 187:4327–4337 Schurr MJ, Deretic V (1997) Microbial pathogenesis in cystic fibrosis: co-ordinate regulation of heat-shock response and conversion to mucoidy in Pseudomonas aeruginosa. Mol Microbiol 24:411–420 Schurr MJ, Martin DW, Mudd MH, Deretic V (1994) Gene cluster controlling conversion of alginate-overproducing phenotype in Pseudomonas aeruginosa: Functional analysis in a heterologous host and role in the instability of mucoidy. J Bacteriol 176:3375–3382 Schurr MJ, Yu H, Boucher JC, Hibler NS, Deretic V (1995) Multiple promoters and induction by heat shock of the gene encoding the alternative sigma factor AlgU (sE) which controls mucoidy in cystic fibrosis isolates of Pseudomonas aeruginosa. J Bacteriol 177:5670–5679 Schurr MJ, Yu H, Martinez-Salazar JM, Boucher JC, Deretic V (1996) Control of AlgU, a member of the sigma E-like family of stress sigma factors, by the negative regulators MucA and MucB and Pseudomonas aeruginosa conversion to mucoidy in cystic fibrosis. J Bacteriol 178:4997–5004 Schwarzmann S, Boring IIIJR (1971) Antiphagocytic effect of slime from a mucoid strain of Pseudomonas aeruginosa. Infect Immun 3:762–767 Seale TW, Thirkhill H, Tarpay M, Flux M, Rennert OM (1979) Serotypes and antibiotic susceptibilities of Pseudomonas aeruginosa isolates from single sputa of cystic fibrosis patients. J Clin Microbiol 9:72–78 Stover CK, X. Q. Pham, A. L. Erwin, S. D. Mizoguchi, P. Warrener, M. J. Hickey, F.S. L. Brinkman, W. O. Hufnagle, D. J. Kowalik, M. Lagrou, R. L. Garber, L. Goltry, E. Tolentino, S. Westbrock-Wadman, Y. Yuan, L. L. Brody, S. N. Coulter, K. R. Folger, A. Kas, K. Larbig, R. Lim, K. Smith, D. Spencer, G. K.-S. Wong, Z. Wu, I. T. Paulsenk, J. Reizer, M. H. Saier, R. E. W. Hancock, Lory S, Olson MV (2000) Complete genome sequence of Pseudomonas aeruginosa PAO1, an opportunistic pathogen. Nature 959:959–964 Suh S-J, Silo-Suh L, Woods D, Hassett D, West S, Ohman D (1999) Effect of rpoS mutation on the stress response and expression of virulence factors in Pseudomonas aeruginosa. J Bacteriol 181:3890–3897 Tart AH, Blanks MJ, Wozniak DJ (2006) The AlgT-dependent transcriptional regulator AmrZ (AlgZ) inhibits flagellum biosynthesis in mucoid, nonmotile Pseudomonas aeruginosa cystic fibrosis isolates. J Bacteriol 188:6483–6489
Alginate Gene Regulation
133
Totten PA, Lara JC, Lory S (1990) The rpoN gene product of Pseudomonas aeruginosa is required for expression of diverse genes, including the flagellin gene. J Bacteriol 172:389–396 Whitchurch CB, Alm RA, Mattick JS (1996) The alginate regulator AlgR and an associated sensor FimS are required for twitching motility in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 93:9839–9843 Whitchurch CB, Erova TE, Emery JA, Sargent JL, Harris JM, Semmler AB, Young MD, Mattick JS, Wozniak DJ (2002) Phosphorylation of the Pseudomonas aeruginosa response regulator AlgR is essential for type IV fimbria-mediated twitching motility. J Bacteriol 184:4544–4554 Wood LF, Ohman DE (2006) Independent regulation of MucD, an HtrA-like protease in Pseudomonas aeruginosa, and the role of its proteolytic motif in alginate gene regulation. J Bacteriol 188:3134–3137 Wood LF, Leech AJ, Ohman DE (2006) Cell wall-inhibitory antibiotics activate the alginate biosynthesis operon in Pseudomonas aeruginosa: roles of sigma (AlgT) and the AlgW and Prc proteases. Mol Microbiol 62:412–426 Woolwine S, Wozniak D (1999) Identification of an Escherichia coli pepA homolog and its involvement in suppression of the algB phenotype in mucoid Pseudomonas aeruginosa. J Bacteriol 181:107–116 Wozniak DJ, Ohman DE (1991) Pseudomonas aeruginosa AlgB, a two-component response regulator of the NtrC family, is required for algD transcription. J Bacteriol 173:1406–1413 Wozniak DJ, Ohman DE (1993) Involvement of the alginate algT gene and integration host factor in the regulation of the Pseudomonas aeruginosa algB gene. J Bacteriol 175:4145–4153 Wozniak DJ, Ohman DE (1994) Transcriptional analysis of the Pseudomonas aeruginosa genes algR, algB, and algD reveals a hierarchy of alginate gene expression which is modulated by algT. J Bacteriol 176:6007–6014 Xie ZD, Hershberger CD, Shankar S, Ye RW, Chakrabarty AM (1996) Sigma factor-anti-sigma factor interaction in alginate synthesis: inhibition of AlgT by MucA. J Bacteriol 178:4990–4996 Yu H, Schurr MJ, Deretic V (1995) Functional equivalence of Escherichia coli sE and Pseudomonas aeruginosa AlgU: E coli rpoE restores mucoidy and reduces sensitivity to reactive oxygen intermediates in algU mutants of P. aeruginosa. J Bacteriol 177:3259–3268 Yu H, Mudd M, Boucher JC, Schurr MJ, Deretic V (1997) Identification of the algZ gene upstream of the response regulator algR and its participation in control of alginate production in Pseudomonas aeruginosa. J Bacteriol 179:187–193
Role of Alginate in Bacterial Biofilms Larry J. Halverson
Contents 1 Introduction ...................................................................................................................... 2 Bacterial Exopolysaccharides .......................................................................................... 3 Alginate Structure and Properties .................................................................................... 4 Regulation of Alginate Biosynthesis ............................................................................... 5 Role of Alginate in Biofilm Formation ............................................................................ 6 Alginate Creates a Hydrated Biofilm Microenvironment ................................................ 7 Alginate Production in the CF Lung ................................................................................ 8 Alginate Protects Cells from Antimicrobial Agents ........................................................ 9 Alginate in Phytopathogenic Interactions ........................................................................ 10 Concluding Remarks and Future Perspectives................................................................. References ................................................................................................................................
136 136 137 138 140 141 143 144 145 147 147
Abstract The prevalence of obtaining alginate overproducing stains from lungs of patients with cystic fibrosis and since alginate is an epiphytic fitness and plant pathogenic virulence trait has promoted inquires into the biological function of alginate. Clues into the role of alginate have been revealed by exploring biofilm matrix composition and alginate biosynthesis regulation at the transcriptional and the posttranslational level. Thus, we are refining our appreciation of the types of environmental stressor that activate alginate production and how surface growth may be an important attribute necessary for alginate production. Alginate production likely occurs under conditions in which cues of environmental stresses and biofilm development processes are integrated into regulatory networks controlling alginate production in a fashion that promotes survival of biofilm residents.
L.J. Halverson Department of Plant Pathology, Iowa State University, Ames, IA 50011, USA e-mail: [email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_6, © Springer-Verlag Berlin Heidelberg 2009
135
136
1
L.J. Halverson
Introduction
Direct observation of a variety of habitats (soil, water, animals, plants) has established that most microbes reside on surfaces to create a structured biofilm ecosystem rather than as planktonic individuals. By definition, biofilms are assemblages of microbial populations encased in an extracellular matrix adhered to each other and/ or to surfaces or interfaces (Costerton et al. 1995). Matrix components can include lipids, proteins, nucleic acids, and/or polysaccharides, such as alginate. Interest in the contribution of alginate to cell adherence to biotic and abiotic surfaces and as the scaffolding material of biofilms led to the development of Pseudomonas aeruginosa as a model for biofilm studies. This focus was due, in part, to the isolation of alginate-overproducing strains from chronic P. aeruginosa infections in the cystic fibrosis (CF) lung (Costerton et al. 1995; Ramsey and Wozniak 2005). Remarkable discoveries have occurred in biofilm research this past decade, particularly on the transition from a motile to a sessile lifestyle, and the production of matrix polysaccharides (Hickman et al. 2005; Lee et al. 2007; Merighi et al. 2007). Moreover, we are gaining new insight into the natural conditions under which bacteria produce alginate and the benefits it confers onto biofilm residents. It is clear from genomic and genetic data that the ability of bacteria to make alginate is widespread amongst members of the ribosomal RNA homology group I of pseudomonads and several Azotobacter species (Cote and Krull 1988; Fett et al. 1992; Gacesa 1998). Alginate plays a key role as a virulence factor of plant-pathogenic pseudomonads, in plant epiphytic fitness, and the encystment process of Azotobacter species in addition to P. aerguinosa infections of CF lungs. Despite the nearly universal ability of Pseudomonas species to produce alginate, we have been unable to determine the exact benefits of alginate production. The role of alginate production in natural environments is not well understood. This is a consequence of difficulties in finding laboratory conditions favoring alginate production and the rapid loss of alginate production by overproducers isolated from CF lungs when they are cultivated in the laboratory. Yet, recent findings are providing new insights into biofilm conditions, the environmental factors, and novel regulatory mechanisms controlling alginate production, in particular in the opportunistic human pathogen P. aeruginosa. From these studies a picture is emerging revealing the central role of alginate in biofilm lifestyle strategies and stress tolerance, in an environmental context- and species-dependent manner that could lead to increased ecological success of its producers.
2
Bacterial Exopolysaccharides
The fact that bacteria produce extracellular polysaccharides is well recognized, largely because of their involvement in both beneficial and detrimental bacteria–host interactions, and environmental stress tolerance. Extracellular polysaccharides have been classified as capsular or exocellular polysaccharides, although the distinction between the two is operational (methodological). In many cases this distinction is based
Role of Alginate in Bacterial Biofilms
137
on how tightly the polymers are associated with the cell. Secreted polysaccharides represent the most extensively studied component of the biofilm matrix and are often assumed to be the most abundant extracellular component. Alginate is usually isolated as an exopolysaccharide, and the extent to which it is more tightly associated with the cell is unclear. Depending upon the species and environmental conditions, the amount and types of exopolysaccharide produced can vary extensively. For a long time it was assumed that alginate was the primary exopolysaccharide matrix component of P. aeruginosa biofilms (Wozniak et al. 2003). It is increasingly clear that alginate is one of many exopolysaccharides that can be made by various Pseudomonas species. Although fluorescent pseudomonads can make other exopolysaccharides, alginate biosynthesis capabilities are the most conserved exopolysaccharide biosynthetic capability on the basis of bioinformatic analysis of sequenced genomes. Hence, alginate production likely plays a fundamentally important role in the ecological success of biofilm residents.
3
Alginate Structure and Properties
Alginate consists of a linear polymer of 1,4-b-linked mannuronic acid and its epimer, guluronic acid. These monomers may be arranged in homopolymeric or heteropolymeric block structures. The mannuronate and guluronate confer anionic characteristics to the exopolysaccharide. P. aerguinosa alginates are normally highly O-acetyated on the second and/or third position of the d-mannuronic acid residue, yielding highly diverse block structures with various degrees of acetylation (Gacesa 1998). The extent of O-acetylation and the block structure define the physiochemical properties of alginate. The high molecular mass of bacterial alginates and its negative charge ensure that the polymer is highly hydrated. Extensive O-acetylation of alginate increases the water-binding capacity of the polymer. See “Material properties of alginates” by Donati and Paoletti in this volume for a more detailed description of the physicochemical properties of alginates. In certain cases, cations such as calcium or magnesium can facilitate gelation by acting as an ionic bridge between polymer strands. Temperature has also been shown to influence gelation, with gels being stabilized at lower temperatures. Therefore, the gelling properties are influenced by the chemistry and structure of the alginate, as well as the physicochemical environment. Alginates which contain polyguluronate form rigid gels in the presence of Ca2+, which may contribute a structural role to the outer cyst cell wall of Azotobacter vinelandii (Gacesa 1998). In contrast, the absence of polyguluronate blocks, such as in P. aeruginosa, produces relatively flexible gels in the presence of Ca2+. The physicochemical properties of alginate depend predominantly on the structure and composition of the alginate. Thus, the water-binding, gel-forming, virulence, and immunogenic properties of alginate will be dependent on controlling the level of acetylation, epimerization, and molecular weight of alginate. At present little is known on how modulation of these chemical properties influences the beneficial attributes of alginate production. For example, it is conceivable that alginates
138
L.J. Halverson
produced under water-limiting conditions will be more acetylated than those produced under water-replete conditions since acetylation is directly associated with the water-holding properties of the polymer.
4
Regulation of Alginate Biosynthesis
Alginate biosynthesis and gene regulation have been under scrutiny for a number of years and are the focus of several chapters in this monograph (See Alginate production: precursor biosynthesis, polymerization and secretion” by Rehm and “Alginate gene regulation” by Ohman in this volume for more information on alginate biosynthesis and gene regulation). Briefly, expression of the alginate biosynthetic operon is under the control of the algD promoter, with the exception of algC, which is also involved in lipopolysaccharide and rhamnolipid synthesis (Zielinski et al. 1991). A key regulatory element controlling transcription from the algD promoter is the alternative sigma factor AlgT, also known as AlgU or s22 (Deretic et al. 1994). Besides controlling alginate production, AlgT is the P. aeruginosa equivalent of the extreme heat-shock sigma factor sE of Escherichia coli (Yu et al. 1995). The E. coli sE signal transduction pathway responds to cell envelope stress and is activated by an intramembrane proteolysis cascade that liberates sequestered cytosolic sE. Congruent with the functions played by the sE of E. coli, AlgT is involved in the regulation of a number of other systems in P. aeruginosa, such as protection from heat shock, osmotic stress, and exogenous oxidants (Firoved and Deretic 2003; Martin et al. 1994; Yu et al. 1995). AlgT activity is repressed by the gene products of the mucABCD locus. One of these genes (mucA) encodes an antisigma factor functioning at the protein level by binding to AlgT (Schurr et al. 1996; Xie et al. 1996). The diversity and the nature of the regulatory elements affecting algD expression reflect the integration of various environmental signals influencing release of sigma AlgT from the negative regulator mucA (Fig. 1). Numerous studies have implicated extreme heat, high osmolarity, desiccation, slow growth rates, oxygen availability, oxidative stress, and cell wall disrupting antibiotics as environmental insults that could disrupt mucA sequestration of cytosolic AlgT. The universality of the signals influencing alginate production in pseudomonads has not been demonstrated, although both P. aerguinosa and P. syringae share some conserved signals, including high concentrations of reactive oxygen species and high osmolarity, for activating transcription from alginate promoters (Keith and Bender 1999). By analogy with other sigma/anti-sigma factor systems, MucA (and MucB) likely controls AlgT activity in response to environmental stimuli, permitting physiological and transient release of AlgT, even in the absence of muc mutations. Future work will need to focus on environmental signals or other regulatory proteins that increase AlgT activity in nonmucoid strains: much of our understanding of alginate gene regulation has been explored in mucA mutant strains. Historically, studies concerning the regulation of alginate biosynthesis have focused on transcriptional control of the algD promoter and algT expression in P. aeruginosa.
Role of Alginate in Bacterial Biofilms
139 Desiccation
Oxidative Stress High salinity Extreme heat
Cell wall disrupting antibiotics Slow growth Others?
Outer membrane Periplasm
Envelope Stress MucC
Disrupt AlgT-MucA(B) interactions Cytoplasmic membrane
MucB
MucD
MucA Alg
T
AlgTreleased
Alginate Regulatory Genes AlgT
Alginate Biosynthesis Operon AlgT
Other AlgT regulon genes AlgT
Fig. 1 Model of environmental stress regulation of alginate biosynthesis. The illustration is a summary of the regulation of alginate gene expression derived from various sources (Deretic et al. 1994; Schnider-Keel et al. 2001; Suh et al. 1999; Wood et al. 2006; Wood and Ohman 2006; Yu et al. 1995)
Recently, it was discovered that one of the alginate biosynthesis proteins, Alg44, contained a PilZ domain, which function as receptors of the secondary signaling molecule cyclic diguanylate (c-di-GMP) (Merighi et al. 2007). These secondary signaling molecules control a number of cellular processes in bacteria by functioning as cofactors for enzymes and regulatory proteins (Galperin 2004). Intracellular concentrations of c-di-GMP are regulated by the opposing activities of diguanylate cyclases that synthesize and phosphodiesterases that degrade the signaling molecule. In general, high c-di-GMP levels promote biofilm development, including formation of multilayer bacterial mats at air–liquid interfaces called “pellicles” (Lee et al. 2007) and exopolysaccharide production (Hickman et al. 2005). Thus, c-di-GMP functions coordinately with transcriptional mechanisms in influencing P. aeruginosa exopolysaccharide production, including alginate (Merighi et al. 2007). Alg44 presumably is a transmembrane protein that functions in facilitating the extrusion of polymerized alginate through the outer membrane pore AlgE, and was shown to bind c-di-GMP. Overexpression of a phosphodiesterase led to a significant reduction
140
L.J. Halverson
in alginate production, while, in contrast, elevated levels of c-di-GMP significantly enhanced alginate production (Merighi et al. 2007). These results demonstrated an important role for c-di-GMP in alginate biosynthesis/export. The question remains whether surface growth is a requirement for c-di-GMP activation of alginate production, much like surface growth is a requirement for activation of the Wsp chemosensory system for P. aeruginosa biofilm formation and Pel exopolysaccharide production (Guvener and Harwood 2007; Lee et al. 2007). Posttranslational control of alginate production may explain previous inconsistencies between algD gene expression at substratum interfaces (Davies et al. 1993) and the apparent lack of alginate in biofilms formed on glass in flow-cell systems.
5
Role of Alginate in Biofilm Formation
Various exopolysaccharides have been shown to be required for the initial attachment of a bacterium to a surface or for the structural development of mature biofilms. The paradigm organism for studying bacterial biofilms is P. aeruginosa, and it was presumed that it produced alginate as the primary exopolysaccharide in biofilms. The first line of evidence supporting an involvement of alginate in adherence to inert surfaces was based on the demonstration that both algC and algD genes are upregulated following adherence of mucoid strains to glass surfaces (Davies et al. 1993; Davies and Geesey 1995). Subsequently, it was demonstrated that alginate overproduction, caused by a mucA mutation, altered biofilm developmental patterns, resulting in biofilms containing more biomass than those formed by the wild type (Hentzer et al. 2001). Biofilms made by alginate overproducers are significantly different from those made by nonalginate overproducing strains, with attached cells of alginate overproducers growing more exclusively as discrete microcolonies, resulting in lower substratum coverage and high structural heterogeneity (Hentzer et al. 2001; Nivens et al. 2001). Absence of alginate O-acetylation substantially decreases surface adherence, thereby delaying biofilm initiation (Nivens et al. 2001). Alginate overproduction by mucA mutants can, in some cases, interfere with biofilm growth on glass surfaces in flow-cell systems (Nivens et al. 2001). For mucoid strain P. aeruginosa FRD1, calcium addition stimulates biofilm thickness by at least tenfold compared with the strain without added calcium. Scanning confocal laser microscopy showed increased spacing between cells for the thick biofilms, and Fourier transform infrared spectroscopy revealed that the material between cells is primarily alginate (Sarkisova et al. 2005). One concern with these finding is that high-level AlgT activity in the mucA mutant could result in expression of genes other than those involved in alginate production and it is those genes that contribute to the altered biofilm phenotype of the alginate overproducer. However, this phenotype was not observed in biofilms formed by an algDmucA double mutant, indicating alginate overproduction is causal to the altered biofilm phenotype. Alginate overproduction leads to a dramatic increase in resistance to the antibiotic tobromycin, which could explain why P. aeruginosa biofilms are so recalcitrant to antibiotic therapy (Wozniak et al. 2003).
Role of Alginate in Bacterial Biofilms
141
The role of alginate in biofilm formation and development in flow-cell systems was challenged by composition analyses of the exopolysaccharide matrix and confocal microscopy in both P. aerguinosa and P. syringae biofilms, which revealed that alginate was not a significant component of the matrix (Laue et al. 2006; Stapper et al. 2004; Wozniak et al. 2003). Not only was alginate not detected in the biofilms, transcriptional reporter gene data indicated that alginate biosynthetic gene expression was not induced on initiation of biofilm formation in a microtiter plate assay. Additionally, Wozniak and colleagues (2003) showed that the structural and antibiotic resistance profiles of wild-type and algD mutant biofilms were indistinguishable. This led to speculation that other polymers, such as DNA or lipopolysaccharide, were the primary scaffolding component. Recent evidence suggests that P. aeruginosa is capable of producing several other exopolysaccharide constituents, a hydrophobic glucose-rich polymer called “Pel” (Friedman and Kolter 2004a, b) and a galactose–mannose-rich polymer called “Psl” (Jackson et al. 2004; Ma et al. 2006, 2007; Matsukawa and Greenberg 2004). Much less is known about these polymers than is known about alginate, but it appears that they contribute to biofilm formation and maintenance in flow-cell systems. However, the role of alginate is unquestionably important in CF pathogenesis. This implies that the onset of biofilm formation in the CF lung during initial colonization likely precludes the switch to a mucoid phenotype (Wozniak et al. 2003). The transition from a nonmucoid to a mucoid phenotype may involve a switch from an unidentified exopolysaccharide, possibly Pel or Psl, which may be the primary exopolysaccharide in non-CF-lung habitats. In light of recent data (Chang et al. 2007), one environmental signal stimulating alginate production in the CF lung could be water starvation.
6
Alginate Creates a Hydrated Biofilm Microenvironment
Bacteria exist in a variety of habitats that are routinely or periodically not saturated with water and residents must integrate cues on water abundance into lifestyle strategies. As in fully hydrated systems, biofilms in unsaturated habitats are encapsulated in an exopolysaccharide layer (Chang and Halverson 2003). Since many exopolysaccharides are hygroscopic, their presence presumably creates a more hydrated microenvironment in the immediate vicinity of the cell, thereby contributing to desiccation tolerance. Despite the long-standing assumption that exopolysaccharide production contributes to desiccation tolerance, there has been relatively little evidence to support the notion that exopolysaccharides, or a particular exopolysaccharide component such as alginate, ameliorate the stresses bacterial cells actually experience under desiccating conditions. Recently, Chang et al. (2007) provided direct evidence that alginate functions to maintain cellular hydration and biofilm formation under desiccating conditions. Moreover, alginate is integral to biofilm architecture under water-limiting conditions but not under water-replete conditions (Chang et al. 2007), which is consistent with reports that alginate does not contribute to biofilm formation in fully hydrated flow-through systems (Wozniak et al. 2003).
142
L.J. Halverson
Chang et al. (2007) assessed the extent to which alginate production by wild-type and algD mutants of P. putida, P. aeruginosa, and P. syringae occurs in response to water limitation. Alginate production increased dramatically under water-limiting but not under high osmolarity conditions, and was regulated at the transcriptional level on the basis of an algD promoter transcriptional fusion assay. Figure 2 shows P. aeruginosa and P. syringae colony biofilms cultivated under water-starvation conditions encapsulated in an alginate-containing matrix. The inability to make alginate decreased exopolysaccharide volume (Fig. 2) and resulted in decreased desiccation tolerance of all three algD mutants compared with their respective wild-type species (Chang et al. 2007). To further understand the role of alginate, they used two complementary approaches to assess whether alginate creates a more hydrated microenvironment that protect cells by slowing the rate of cellular drying. Using P. putida as a model, they demonstrated that cells encapsulated with alginate perceive less water stress than those without alginate following a desiccation stress, as measured by an intracellular water potential biosensor (Axtell and Beattie 2002). Second, alginate decreased the extent of membrane dehydration, given that the wild type exhibited fewer membrane fatty acid changes reflective of desiccation-mediated membrane stress compared with the algD mutant. Collectively, the results indicate that alginate retains water in the cell microenvironment either directly, owing to its hygroscopic properties, or through its ability to influence biofilm architecture that
Fig. 2 Morphology of 2-day-old Pseudomonas aeruginosa PAO1 and P. syringae pv. syringae (Pss) B728a colony biofilms under water-limiting conditions. Water starvation stimulates a mucoid PAO1 wild type and PssB728a wild type, a hypermucoid PDO300 mucA, and a nonmucoid PAO1 algD and PssB728a algD colony morphology. Water starvation was imposed by poly(ethylene glycol) amendments to tryptone yeast extract (TYE) medium to lower the water potential by 1.5 MPa as described previously (Chang and Halverson 2003). All strains have a nonmucoid colony morphology on TYE without poly(ethylene glycol) (not shown)
Role of Alginate in Bacterial Biofilms
143
reduces evaporative water loss, or both. Significantly, their findings suggest that the prevalence of alginate biosynthesis capabilities among pseudomonads is because alginate production capabilities are an important fitness trait in soil, the rhizosphere, aerial leaf surface, or the CF lung that can, at times, be water-limited.
7
Alginate Production in the CF Lung
P. aeruginosa infections of lungs of CF patients initially occur as nonmucoid variants. As the P. aeruginosa population expands, it diversifies, yielding strains with traits uncharacteristic of environmental isolates. The same traits are consistently acquired during chronic infection of different patients, suggesting that there is a conserved pattern of evolution by which this opportunistic pathogen adapts to the CF airway (Deretic et al. 1994). The phenotypes acquired by P. aeruginosa over the course of CF infections reflect alterations in diverse aspects of their biology. One commonly isolated variant is for those that exhibit a mucoid colony morphology, due to the overproduction of alginate. Thus, alginate is presumed to protect P. aeruginosa in the CF lung. Once the infecting bacteria start overproducing alginate, the infection is very difficult to eradicate and is associated with a poor prognosis (Burns et al. 2001; Ciofu et al. 2008; Singh et al. 2000). An important visual observation of CF sputum samples suggested that P. aerguinosa in the CF lung forms biofilms (microcolonies) consisting of exopolysaccharide-embedded cells (Deretic et al. 1994; Singh et al. 2000). Studies on CF-derived mucoid P. aeruginosa isolates in early biofilm studies led to the assumption that alginate is the primary exopolysaccharide in biofilms formed by mucoid and nonmucoid isolates, although this perspective is changing. Inhibition of adherence of nonmucoid but not mucoid strains to eukaryotic cells grown in vitro by alginate antibodies was also considered evidence for the role of alginate in biofilm formation. Consequently, the role of alginate in the formation of P. aeruginosa biofilms has been the focus of numerous studies, as described earlier in this chapter, leading to the conclusion that alginate is not a component of the biofilm matrix in flow-cell systems (Wozniak et al. 2003). The question arises as to the identity of the environmental parameters that exert such a strong selective pressure favoring alginate-overproducing variants in the CF lung. Mucoidy arises owing to mutations in one of several different negative regulators (usually mucA) of the alternative sigma factor AlgT, leading to uncontrolled AlgT activation of the alginate biosynthesis genes, as described earlier. Consequently, these variants overproduce alginate, as illustrated in Fig. 2, where the normally mucoid mucA mutant strain PDO300 becomes hypermucoid compared with its wild-type parent under water-limiting conditions. Water starvation stimulates the AlgT-mediated envelope stress response in pseudomonads (van de Mortel and Halverson 2004). Besides controlling alginate production, AlgT also controls a large repertoire of other genes involved in stress tolerance (Yu et al. 1995). Since the presence of uncontrolled AlgT activity in mucoid isolates interferes with normal cellular physiological function, the selective advantage of mucoid forms must be
144
L.J. Halverson
critical for the ability of the organisms to persistently infect CF patients. Alternatively, it is the altered physiological function (Firoved and Deretic 2003) and not alginate production per se that provides a competitive advantage in the CF lung. The metabolic costs associated with alginate production must be great since mucoid strains cultivated in the laboratory rapidly lose alginate-production capabilities and many CF isolates have mutations in mucA and in algT, resulting in nonmucoid variants. There is an increasing body of evidence indicating that the CF lung environment is water-limited. The periciliary liquid layer depletion model suggests that CF airway epithelia exhibit abnormally high rates of airway surface liquid absorption (Mall et al. 2004; Matsui et al. 1998, 2005). Presumably, loss of water volume from the CF epithelia results in thickening of mucus, which restricts its transport and provides a favorable environment for microbial colonization (Mall et al. 2004; Matsui et al. 2006). Mucus production may counter water loss if the mucus water sorption properties are sufficient to counter the high rate of liquid adsorption. Biophysical measurements suggest mucus is sufficiently dehydrated to restrict bacterial motility and small-molecule diffusion (Matsui et al. 2005, 2006). In light of recent reports indicating that water deprivation stimulates alginate production (Chang et al. 2007), selective pressure for variants with high AlgT activity, such as mucA variants, may be particularly strong in the CF lung environment.
8
Alginate Protects Cells from Antimicrobial Agents
Alginate is clearly produced in the CF lung since antibodies to P. aerguinosa alginate have been detected in CF patients, although no mucoid cultures were isolated from sputum samples (Pedersen et al. 1990). With use of alginate-specific antibodies it was also shown that shortly after infection of mice alginate is expressed (Pier et al. 2004). Additionally, alginate gene expression has been observed in nonmucoid P. aeruginosa in the lungs of patients with CF and in a mouse model (Bragonzi et al. 2005). Evidence indicates that CF lungs are initially colonized by nonmucoid strains, followed by the appearance of mucoid variants. Alginate does provide protection to P. aeruginosa in the CF lung. Infection of the CF lung causes inflammatory cells to be recruited to the site of infection, where they release reactive oxygen species to kill the invading microbes and, unfortunately, cause extensive tissue damage (Ramsey and Wozniak 2005). Alginate can suppress leukocyte functions and nonopsonic phagocytosis and promote adhesion to eukaryotic cell surfaces (Mai et al. 1993a, b; Ramphal and Pier 1985). Complementing this effect is the reported ability of alginate to scavenge reactive oxygen species such as superoxide radicals and hypochlorite, which are likely to play an important role during in vivo killing by macrophages and neutrophils (Learn et al. 1987; Simpson et al. 1989). Furthermore, increased AlgT activity also causes increased expression of antioxidant systems in P. aeruginosa, leading to additional protection from reactive oxygen species (Firoved and Deretic 2003; Firoved et al. 2004). Alginate also protects P. aeruginosa cells from interferon-g-mediated macrophage
Role of Alginate in Bacterial Biofilms
145
killing (Leid et al. 2005). Alginate probably provides a physical and chemical barrier, thereby protecting bacteria from phagocytosis. There has been much speculation that alginate production by biofilm cells increases resistances to antibiotics, although the mode of growth may be more important than alginate production in resisting antibiotics. It is difficult to ascertain the extent to which alginate overproduction contributes to increased antibiotic resistance since increased AlgT activity may induce genes other than those for alginate production that confer increased antibiotic resistance properties. However, it was shown that in nonmucoid P. aerguinosa biofilms exposed to subinhibitory levels of the b-lactam antibiotic imipenem, alginate gene transcription was upregulated and coincided with increased alginate production and biofilm volume (Bagge et al. 2004). This observation was subsequently verified to demonstrate that cell-wall inhibitory antibiotics activate alginate biosynthesis and that the AlgW and Prc proteases play a fundamentally important role (Wood et al. 2006). These observations are consistent with cell-wall inhibiting antibiotics in stimulating the envelope stress response, presumably because disruption of the cell wall leads to disruption of the AlgT-MucA signal transduction system. In contrast, the macrolide antibiotic azithromycin blocks quorum-sensing and alginate production, which increases cell sensitivity to complement (Hoffmann et al. 2007). This may explain why azithromycin is clinically effective.
9
Alginate in Phytopathogenic Interactions
The virulence of numerous phytopathogenic bacteria has been correlated with their ability to produce exopolysaccharides. Several studies have shown that the major exopolysaccharide produced by P. syringae in planta is alginate in water-soaked leaves (Fett and Dunn 1989). P. syringae can also produce the polysaccharide levan, although it is less clear whether it is made in planta. The infection of host plants by P. syringae involves epiphytic (surface) colonization, entry, establishment of infection sites in the intercellular spaces (apoplast), multiplication within host tissue, and production of disease symptoms. During infection of nonhost plants, the hypersensitive response (HR) functions to restrict pathogen growth and spread. Alginate production has also been associated with increased epiphytic fitness and resistance to toxic molecules (Keith and Bender 1999; Kidambi et al. 1995; Yu et al. 1999). The aerial surface of plants can be an extremely harsh environment, with high levels of solar radiation and fluctuating cycles of wetness. It is likely a suite of adaptations are necessary for epiphytes to survive the harsh leaf environment, and alginate may play a role. Disruption of alginate lyase (algL) by transposon mutagenesis revealed that an alginate-deficient P. syringae pv. syringae mutant was unable to colonize the leaf surface of nonhost plants as well as the wild type (Yu et al. 1999). Evaluation of epiphytic fitness was based on survival on tomato leaves that were kept dry in a low-humidity greenhouse environment. This suggests
146
L.J. Halverson
that alginate was particularly important for colonizing dry leaf surfaces. However, lesions formed by the alginate-deficient mutant were less severe, possibly because of their lower in planta populations (Yu et al. 1999). Collectively, these results suggested that alginate contributes to virulence by facilitating colonization or dissemination of the bacterium in planta (Yu et al. 1999). In other words, higher population sizes on leaf surfaces were a consequence of better in planta survival. Although alginate has been isolated from plants infected by P. syringae, the signals and timing of alginate gene expression in planta have not been fully described. On the basis of various reporter gene studies, the alginate biosynthesis operon is expressed in susceptible plants (Keith et al. 2003). Significantly, algD gene expression occurred more quickly and to a greater extent in resistant than in susceptible tomato ‘Rio Grande’ plants (Keith et al. 2003). Expression of algD was evident in individuals colonizing the leaf surface, suggesting biofilm formation is not a prerequisite for alginate production. Within 12h after inoculation, microscopic HRwas visible in resistant tomato cultivars and the incompatible host tobacco inoculated with P. syringae pv. tomato. Yet, expression of algD was visible within 8–12h after inoculation. When the incompatible host tobacco was inoculated with an hrpC mutant of P. syringae, the HR did not occur and algD expression was substantially lower. hrpC mutants do not elicit a HR in tobacco (Keith et al. 2003). Collectively the data suggest that signals that proceed the HR may stimulate alginate gene expression in P. syringae. One possible signal for algD expression is superoxide anion since the oxidative burst, which involves the production of potentially cytotoxic quantities of hydrogen peroxide and superoxide anion, is associated with the HR. The oxidative burst may occur before the HR is visible and reactive oxygen species can induce alginate gene expression in P. syringae (Yu et al. 1999). Another possible signal is water stress associated with the HR. The HR is a complex form of programmed cell death that is generally characterized by the presence of brown, dead plant cells at the infection site. The HR is also associated with cessation of the growth and spread of the bacteria. Interestingly, evidence of alginate gene expression by P. syringae pv. tomato in planta (Yu et al. 1999) preceding the HR coincides with inhibitory levels of water stress during the HR (Wright and Beattie 2004). Wright and Beattie used a transcriptional fusion between the E. coli proU promoter that responds to both the solute and matric components of the total water potential and responds specifically to low water potential and not to ion toxicity or oxidative stress. By measuring the water potential that virulent and avirulent P. syringae encounter during their interactions with Arabidopsis thaliana, they observed that water potential decreases occurred within 4h after inoculation of the plants and well before visible signs of the HR (Wright and Beattie 2004). The water potential sensed by the bacteria during pathogenesis was sufficient to prevent cell division in the majority of cells in culture. It is conceivable that in addition to or coincident with reactive oxygen species mediated induction of alginate synthesis in planta, water starvation occurs during the initial stages of the HR. Alginate may be responsible for the water-soaked appearance of lesions given it is highly hydrophilic and would help maintain moisture in the apoplast to offset HR-mediated desiccation of the infection site.
Role of Alginate in Bacterial Biofilms
10
147
Concluding Remarks and Future Perspectives
Research in the last few years has shed new light onto the potential function of alginate production in bacterial biofilm development and environmental stress tolerance. Much research still has to be done to further define the environmental conditions under which alginate production occurs. Clearly, alginate production is controlled by factors that stimulate the envelope stress response, such as water starvation (van de Mortel and Halverson 2004) and cell-wall inhibiting antibiotics (Wood et al. 2006). Other environmental stressors may stimulate alginate production in a habitat- and species-dependent manner. The recent report that c-di-GMP functions to control alginate production at the posttranslational level provides a foundation on which to further refine our understanding of how this regulatory mechanism influences alginate production. For example, is surface growth a perquisite for c-di-GMP activation of alginate production by P. aeruginosa much like it is for production of the hydrophobic Pel polysaccharide? How is alginate production coordinated with the production of other exopolysacchrides that are components of the biofilm matrix since alginate may be only a minor component of the matrix of nonmucoid strains? Of particular interest will be to reveal the spatiotemporal patterns of alginate production within biofilms, particularly since alginate production is energetically demanding under conditions amendable to activating the envelope stress response. Since exopolysaccharides are a shared resource, alginate production by certain members of the microbial community may provide protection to non-alginate-producing residents. If this is the case it could explain, in part, the fact that not all P. aeruginosa strains isolated from the CF lung environment overproduce alginate and why some have lost the ability to make alginate. The prevalence of alginate biosynthesis abilities amongst pseudomonads implies that alginate is an important fitness trait under certain environmental conditions or in particular habitats. Understanding the competitive advantages of producing alginate by biofilm residents will provide great insight into the biology of various host-pathogen interactions. Acknowledgements The author is indebted to the many thoughtful discussions with colleagues that contributed to the ideas expressed in this chapter. I also acknowledge the financial support provided by the US National Science Foundation and Department of Agriculture to pursue our interest in how water availability influences bacterial alginate production.
References Axtell CA, Beattie GA (2002) Construction and characterization of a proU-gfp transcriptional fusion that measures water availability in a microbial habitat. Appl Environ Microbiol 68: 4604–4612 Bagge N, Schuster M, Hentzer M, Ciofu O, Givskov M, Greenberg EP, Hoiby N (2004) Pseudomonas aeruginosa biofilms exposed to imipenem exhibit changes in global gene expression and beta-lactamase and alginate production. Antimicrob Agents Chemother 48:1175–1187
148
L.J. Halverson
Bragonzi A, Worlitzsch D, Pier GB, Timpert P, Ulrich M, Hentzer M, Andersen JB, Givskov M, Conese M, Doring G (2005) Nonmucoid Pseudomonas aeruginosa expresses alginate in the lungs of patients with cystic fibrosis and in a mouse model. J Infect Dis 192:410–419 Burns JL, Gibson RL, McNamara S, Yim D, Emerson J, Rosenfeld M, Hiatt P, McCoy K, Castile R, Smith AL, Ramsey BW (2001) Longitudinal assessment of Pseudomonas aeruginosa in young children with cystic fibrosis. J Infect Dis 183:444–452 Chang W-S, Halverson LJ (2003) Reduced water availability influences the dynamics, development, and ultrastructural properties of Pseudomonas putida biofilms. J Bacteriol 185:6199–6204 Chang WS, van de Mortel M, Nielsen L, Nino de Guzman G, Li X, Halverson LJ (2007) Alginate production by Pseudomonas putida creates a hydrated microenvironment and contributes to biofilm architecture and stress tolerance under water-limiting conditions. J Bacteriol 189:8290–8299 Ciofu O, Lee B, Johannesson M, Hermansen NO, Meyer P, Hoiby N, The Scandinavian Cystic Fibrosis Study C (2008) Investigation of the algT operon sequence in mucoid and non-mucoid Pseudomonas aeruginosa isolates from 115 Scandinavian patients with cystic fibrosis and in 88 in vitro non-mucoid revertants. Microbiology 154:103–113 Costerton JW, Lewandowski Z, Caldwell DE, Korber DR, Lappin-Scott HM (1995) Microbial biofilms. Annu Rev Microbiol 49:711–745 Cote GL, Krull LH (1988) Characterization of the exocellular polysaccharides from Azotobacter chroococcum. Carbohydr Res 181:143–152 Davies DG, Geesey GG (1995) Regulation of the alginate biosynthesis gene algC in Pseudomonas aeruginosa during biofilm development in continuous culture. Appl Environ Microbiol 61:860–867 Davies DG, Chakrabarty AM, Geesey GG (1993) Exopolysaccharide production in biofilms: substratum activation of alginate gene expression by Pseudomonas aeruginosa. Appl Environ Microbiol 59:1181–1186 Deretic V, Schurr MJ, Boucher JC, Martin DW (1994) Conversion of Pseudomonas aeruginosa to mucoidy in cystic fibrosis: environmental stress and regulation of bacterial virulence by alternative sigma factors. J Bacteriol 176:2773–2780 Fett WF, Dunn MF (1989) Exopolysaccharides produced by phytopathogenic Pseudomonas syringae pathovars in infected leaves of susceptible hosts. Plant Physiol 89:5–9 Fett WF, Wijey C, Lifson ER (1992) Occurrence of alginate gene sequences among members of the pseudomonad rRNA homology groups I-IV. FEMS Microbiol Lett 78:151–157 Firoved AM, Deretic V (2003) Microarray analysis of global gene expression in mucoid Pseudomonas aeruginosa. J Bacteriol 185:1071–1081 Firoved AM, Wood SR, Ornatowski W, Deretic V, Timmins GS (2004) Microarray analysis and functional characterization of the nitrosative stress response in nonmucoid and mucoid Pseudomonas aeruginosa. J Bacteriol 186:4046–4050 Friedman L, Kolter R (2004a) Genes involved in matrix formation in Pseudomonas aeruginosa PA14 biofilms. Mol Microbiol 51:675–690 Friedman L, Kolter R (2004b) Two genetic loci produce distinct carbohydrate-rich structural components of the Pseudomonas aeruginosa biofilm matrix. J Bacteriol 186:4457–4465 Gacesa P (1998) Bacterial alginate biosynthesis - recent progress and future prospects. Microbiology 144:1133–1143 Galperin MY (2004) Bacterial signal transduction network in a genomic perspective. Environ Microbiol 6:552–567 Guvener ZT, Harwood CS (2007) Subcellular location characteristics of the Pseudomonas aeruginosa GGDEF protein, WspR, indicate that it produces cyclic-di-GMP in response to growth on surfaces. Mol Microbiol 66:1459–1473 Hentzer M, Teitzel GM, Balzer GJ, Heydorn A, Molin S, Givskov M, Parsek MR (2001) Alginate overproduction affects Pseudomonas aeruginosa biofilm structure and function. J Bacteriol 183:5395–5401 Hickman JW, Tifrea DF, Harwood CS (2005) A chemosensory system that regulates biofilm formation through modulation of cyclic diguanylate levels. Proc Nat Acad Sci U S A 102:14422–14427
Role of Alginate in Bacterial Biofilms
149
Hoffmann N, Lee B, Hentzer M, Rasmussen TB, Song Z, Johansen HK, Givskov M, Hoiby N (2007) Azithromycin blocks quorum sensing and alginate polymer formation and increases the sensitivity to serum and stationary-growth-phase killing of Pseudomonas aeruginosa and attenuates chronic P. aeruginosa lung infection in Cftr(–/–) mice. Antimicrob Agents Chemother 51:3677–3687 Jackson KD, Starkey M, Kremer S, Parsek MR, Wozniak DJ (2004) Identification of psl, a locus encoding a potential exopolysaccharide that is essential for Pseudomonas aeruginosa PAO1 biofilm formation. J Bacteriol 186:4466–4475 Keith LMW, Bender CL (1999) AlgT (s22) controls alginate production and tolerance to environmental stress in Pseudomonas syringae. J Bacteriol 181:7176–7184 Keith RC, Keith LMW, Hernandez-Guzman G, Uppalapati SR, Bender CL (2003) Alginate gene expression by Pseudomonas syringae pv. tomato DC3000 in host and non-host plants. Microbiology 149:1127–1138 Kidambi SP, Sundin GW, Palmer DA, Chakrabarty AM, Bender CL (1995) Copper as a signal for alginate synthesis in Pseudomonas syringae pv. syringae. Appl Environ Microbiol 61:2172–2179 Laue H, Schenk A, Li H, Lambertsen L, Neu TR, Molin S, Ullrich MS (2006) Contribution of alginate and levan production to biofilm formation by Pseudomonas syringae. Microbiology 152:2909–2918 Learn DB, Brestel EP, Seetharama S (1987) Hypochlorite scavenging by Pseudomonas aeruginosa alginate. Infect Immun 55:1813–1818 Lee VT, Matewish JM, Kessler JL, Hyodo M, Hayakawa Y, Lory S (2007) A cyclic-di-GMP receptor required for bacterial exopolysaccharide production. Mol Microbiol 65:1474–1484 Leid JG, Willson CJ, Shirtliff ME, Hassett DJ, Parsek MR, Jeffers AK (2005) The exopolysaccharide alginate protects Pseudomonas aeruginosa biofilm bacteria from IFN-gamma-mediated macrophage killing. J Immunol 175:7512–7518 Ma L, Jackson KD, Landry RM, Parsek MR, Wozniak DJ (2006) Analysis of Pseudomonas aeruginosa conditional Psl variants reveals roles for the Psl polysaccharide in adhesion and maintaining biofilm structure postattachment. J Bacteriol 188:8213–8221 Ma L, Lu H, Sprinkle A, Parsek MR, Wozniak DJ (2007) Pseudomonas aeruginosa Psl is a galactose- and mannose-rich exopolysaccharide. J Bacteriol 189:8353–8356 Mai GT, McCormack JG, Seow WK, Pier GB, Jackson LA, Thong YH (1993a) Inhibition of adherence of mucoid Pseudomonas aeruginosa by alginase, specific monoclonal antibodies, and antibiotics. Infect Immun 61:4338–4343 Mai GT, Seow WK, Pier GB, McCormack JG, Thong YH (1993b) Suppression of lymphocyte and neutrophil functions by Pseudomonas aeruginosa mucoid exopolysaccharide (alginate): reversal by physicochemical, alginase, and specific monoclonal antibody treatments. Infect Immun 61:559–564 Mall M, Grubb BR, Harkema JR, O’Neal WK, Boucher RC (2004) Increased airway epithelial Na+ absorption produces cystic fibrosis-like lung disease in mice. Nat Med 10:487–493 Martin DW, Schurr MJ, Yu H, Deretic V (1994) Analysis of promoters controlled by the putative sigma factor AlgU regulating conversion to mucoidy in Pseudomonas aeruginosa: relationship to sE and stress response. J Bacteriol 176:6688–6696 Matsui H, Grubb BR, Tarran R, Randell SH, Gatzy JT, Davis CW, Boucher RC (1998) Evidence for periciliary liquid layer depletion, not abnormal ion composition, in the pathogenesis of cystic fibrosis airways disease. Cell 95:1005–1015 Matsui H, Verghese MW, Kesimer M, Schwab UE, Randell SH, Sheehan JK, Grubb BR, Boucher RC (2005) Reduced three-dimensional motility in dehydrated airway mucus prevents neutrophil capture and killing bacteria on airway epithelial surfaces. J Immunol 175:1090–1099 Matsui H, Wagner VE, Hill DB, Schwab UE, Rogers TD, Button B, Taylor RM, II, Superfine R, Rubinstein M, Iglewski BH, Boucher RC (2006) A physical linkage between cystic fibrosis airway surface dehydration and Pseudomonas aeruginosa biofilms. Proc Nat Acad Sci U S A 103:18131–18136
150
L.J. Halverson
Matsukawa M, Greenberg EP (2004) Putative exopolysaccharide synthesis genes influence Pseudomonas aeruginosa biofilm development. J Bacteriol 186:4449–4456 Merighi M, Lee VT, Hyodo M, Hayakawa Y, Lory S (2007) The second messenger bis-(3¢-5¢)cyclic-GMP and its PilZ domain-containing receptor Alg44 are required for alginate biosynthesis in Pseudomonas aeruginosa. Mol Microbiol 65:876–895 Nivens DE, Ohman DE, Williams J, Franklin MJ (2001) Role of alginate and its O acetylation in formation of Pseudomonas aeruginosa microcolonies and biofilms. J Bacteriol 183:1047–1057 Pedersen SS, Espersen F, Hoiby N, Jensen T (1990) Immunoglobulin A and immunoglobulin G antibody responses to alginates from Pseudomonas aeruginosa in patients with cystic fibrosis. J Clin Microbiol 28:747–755 Pier GB, Boyer D, Preston M, Coleman FT, Llosa N, Mueschenborn-Koglin S, Theilacker C, Goldenberg H, Uchin J, Priebe GP, Grout M, Posner M, Cavacini L (2004) Human monoclonal antibodies to Pseudomonas aeruginosa alginate that protect against infection by both mucoid and nonmucoid strains. J Immunol 173:5671–5678 Ramphal R, Pier GB (1985) Role of Pseudomonas aeruginosa mucoid exopolysaccharide in adherence to tracheal cells. Infect Immun 47:1–4 Ramsey DM, Wozniak DJ (2005) Understanding the control of Pseudomonas aeruginosa alginate synthesis and the prospects for management of chronic infections in cystic fibrosis. Mol Microbiol 56:309–322 Sarkisova S, Patrauchan MA, Berglund D, Nivens DE, Franklin MJ (2005) Calcium-induced virulence factors associated with the extracellular matrix of mucoid Pseudomonas aeruginosa biofilms. J Bacteriol 187:4327–4337 Schnider-Keel U, Lejbølle KB, Baehler E, Haas D, Keel C (2001) The sigma factor AlgU (AlgT) controls exopolysaccharide production and tolerance towards desiccation and osmotic stress in the biocontrol agent Pseudomonas fluorescens CHA0. Appl Environ Microbiol 67:5683–5693 Schurr MJ, Yu H, Martínez-Salazar JM, Boucher JC, Deretic V (1996) Control of AlgU, a member of the sE-like family of stress sigma factors, by the negative regulators MucA and MucB and Pseudomonas aeruginosa conversion to mucoidy in cystic fibrosis. J Bacteriol 178:4997–5004 Simpson JA, Smith SE, Dean RT (1989) Scavenging by alginate of free radicals released by macrophages. Free Radic Biol Med 6:347–353 Singh PK, Schaefer AL, Parsek MR, Moninger TO, Welsh MJ, Greenberg EP (2000) Quorum-sensing signals indicate that cystic fibrosis lungs are infected with bacterial biofilms. Nature 407:762–764 Stapper AP, Narasimhan G, Ohman DE, Barakat J, Hentzer M, Molin S, Kharazmi A, Høiby N, Mathee K (2004) Alginate production affects Pseudomonas aeruginosa biofilm development and architecture, but is not essential for biofilm formation. J Med Microbiol 53:679–690 Suh S-J, Silo-Suh L, Woods DE, Hassett DJ, West SE, Ohman DE (1999) Effect of rpoS mutation on the stress response and expression of virulence factors in Pseudomonas aeruginosa. J Bacteriol 181:3890–3897 van de Mortel M, Halverson LJ (2004) Cell envelope components contributing to biofilm growth and survival of Pseudomonas putida in low-water-content habitats. Mol Microbiol 52:735–750 Wood LF, Ohman DE (2006) Independent regulation of MucD, an HtrA-like protease in Pseudomonas aeruginosa, and the role of its proteolytic motif in alginate gene regulation. J Bacteriol 188:3134–3137 Wood LF, Leech AJ, Ohman DE (2006) Cell wall-inhibitory antibiotics activate the alginate biosynthesis operon in Pseudomonas aeruginosa: roles of sigma (AlgT) and the AlgW and Prc proteases. Mol Microbiol 62:412–426 Wozniak DJ, Wyckoff TJO, Starkey M, Keyser R, Azadi P, O’Toole GA, Parsek MR (2003) Alginate is not a significant component of the extracellular polysaccharide matrix of PA14 and PAO1 Pseudomonas aeruginosa biofilms. Proc Natl Acad Sci U S A 100:7907–7912
Role of Alginate in Bacterial Biofilms
151
Wright CA, Beattie GA (2004) Pseudomonas syringae pv. tomato cells encounter inhibitory levels of water stress during the hypersensitive response of Arabidopsis thaliana. Proc Natl Acad Sci U S A 101:3269–3274 Xie ZD, Hershberger CD, Shankar S, Ye RW, Chakrabarty AM (1996) Sigma factor-anti-sigma factor interaction in alginate synthesis: inhibition of AlgT by MucA. J Bacteriol 178:4990–4996 Yu H, Schurr MJ, Deretic V (1995) Functional equivalence of Escherichia coli sE and Pseudomonas aeruginosa AlgU: E. coli rpoE restores mucoidy and reduces sensitivity to reactive oxygen intermediates in algU mutants of P. aeruginosa. J Bacteriol 177:3259–3268 Yu J, Peñaloza-Vázquez A, Chakrabarty AM, Bender CL (1999) Involvement of the exopolysaccharide alginate in the virulence and epiphytic fitness of Pseudomonas syringae pv. syringae. Mol Microbiol 33:712–720 Zielinski NA, Chakrabarty AM, Berry A (1991) Characterization and regulation of the Pseudomonas aeruginosa algC gene encoding phosphomannomutase. J Biol Chem 266:9754–9763
Microbial Production of Alginates: Physiology and Process Aspects Wael Sabra and An Ping Zeng
Contents 1 2 3
Introduction........................................................................................................................ Commercial Production of Alginate .................................................................................. Biological Role of Alginate Production in Bacteria .......................................................... 3.1 Azotobacter vinelandii .............................................................................................. 3.2 Pseudomonads .......................................................................................................... 4 Optimal Conditions for Microbial Alginate Production .................................................... 4.1 Fermentation Conditions for Optimal Alginate Production in A. vinelandii ........... 4.2 Alginate Production by P. aeruginosa ....................................................................... 4.3 Kinetics of Alginate Production ............................................................................... 4.4 Non-Newtonian Alginate Fluids and the Role of Fermentor Hydrodynamics ......... 5 Conclusion ......................................................................................................................... References ................................................................................................................................
154 154 156 156 159 162 162 164 165 166 168 168
Abstract Presently, most of the alginate produced commercially is still obtained from algae, although it is subjected to variations in quality and quantity due to changes of climate and sources. Alginate can also be biotechnologically produced by species of two families of heterotrophic bacteria, Pseudomonas and Azotobacter. Efforts have been made in the past to produce alginate-like polymers from these bacteria. The association of virulence with alginate production in most Pseudomonas spp. has made Azotobacter vinelandii the most promising candidate for the industrial production of alginate. Nevertheless, for specific and well-defined applications, especially in biomedical and pharmaceutical fields, the production of Pseudomonas alginate has attracted increasing attention. Microbial alginate production has been widely investigated in batch, fed-batch, and continuous cultures. This chapter summarizes current knowledge of physiology and process aspects in view of potential industrial production of microbial alginate. W. Sabra Botany and Microbiology Department, Faculty of Science, Alexandria University, Alexandria, Egypt A.P. Zeng() Bioprocess and Biosystems Engineering, Hamburg University of Technology, Hamburg University of Technology (TUHH), Denickerstrasse 15, 21073, Hamburg, Germany e-mail: [email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_7, © Springer-Verlag Berlin Heidelberg 2009
153
154
1
W. Sabra and A.P. Zeng
Introduction
As early as 600 b.c., seaweed was used as food for man, but algin, a component of seaweed, was only first discovered in 1880 by the British chemist Stanford. Pure alginic acid was prepared in 1896 by Akrefting. In 1929 the company Kelco began commercial production of alginate and introduced milk-soluble alginate as an ice cream stabilizer. The unique physical properties of alginate enable it to be used as a stabilizer, viscosifier, and gelling agent in the food and beverage, paper and printing, biomaterials, and pharmaceutical industries. However, the high production costs or price of these products, and the environmental impacts associated with seaweed harvesting and processing, have prevented a major breakthrough in various applications. In the middle of the twentieth century, an interesting possible commercial avenue for a microbial polysaccharide was paved by the discovery that strains of Pseudomonas aeruginosa (Linker and Jones 1966) and Azotobacter vinelandii (Gorin and Spencer 1966) secrete extracellular polyuronides that closely resemble alginic acid recovered from brown algae. In the meantime, the interest in microbial alginate has also been very much medically oriented because of its association with the pathogenicity in strains of P. aeruginosa. The fast development of pharmaceutical applications of this polymer as well as the discovery of its unique immunological properties aroused the interest of industrial biochemists in developing a microbially optimized production process of this useful compound. Microbial alginate production has been investigated to date in batch, fed-batch, and continuous cultures. In this chapter we first give a short overview of commercial production of alginate and then summarize several aspects of physiology, such as the biological roles of alginate production in bacteria and bioprocessing, such as optimal conditions for microbial alginate production in bioreactors.
2
Commercial Production of Alginate
The extraction of alginate from algal material is illustrated schematically in Fig. 1. In the first step, acidification of insoluble counterion-alginate (Na+, Mg2+, Ca2+, Sr2+, etc. via ion-exchange equilibrium with the seawater) is done by extracting the milled algal tissue with 0.1–0.2M mineral acid. Alginic acid is then brought into solution by neutralization with an alkali such as sodium carbonate or sodium hydroxide to from water-soluble sodium alginate. Removal of other algal materials is done by various separation methods, such as sifting, flotation, centrifugation, and filtration. Subsequently, sodium alginate is precipitated by addition of alcohol, calcium chloride, or mineral acid, reconverted to the sodium form if needed, and finally dried and milled. About 25 years ago, Young (1983) suggested a similar route for microbial production of alginate, as shown in Fig. 2. It should be emphasized that to maximize the microbial alginate production on a commercial scale, the biological role of alginate formation should be well understood. In fact there are still many questions remaining to be answered in this respect.
Microbial Production of Alginates: Physiology and Process Aspects Alkali +water + heat
Water Wet or dry seaweed
Milling
155
Washing
Washed seaweed
Dissolution of alginates Clarification Insoluble residue
Washings
Crude alginate solution Alginate powder Water and Dissolved impurities
Precipitation
CaCl2
Ca-alginate Drying-milling
NaCO3 incorporation
Na-alginate
Acid treatment and, precipitation Acid
Fig. 1 The manufacture of sodium alginate from seaweed
Medium and alkali
pH control Temperature control Laboratory work
Sterile air Inoculum fermentor
Production fermentor
CaCl2 Acid, water
Cell removal Precipitator and separator Contocter and separator Contocter and separator cells Dewater, dry, mill waste NaCO3
Package
Fig. 2 The production of the microbial alginate (as proposed by Young 1983)
Holding tank
156
3
W. Sabra and A.P. Zeng
Biological Role of Alginate Production in Bacteria
In natural environment in which microorganisms capable of producing exopolysaccharides (EPS) are found, the polysaccharides may be associated with virulence as in the case of plant or animal pathogens, or protect the microbial cell against desiccation and attack by bacteriophages and protozoa or act against macrophages as in the case of human pathogens. The polymers can vary considerably in their chemical structures, and some microorganisms can synthesize more than one type of EPS in accordance with different environmental stimuli.
3.1
Azotobacter vinelandii
3.1.1
Alginate Synthesis Is Essential for Cyst Formation
A. vinelandii is distinguished from the rest of the family Azotobacteriaceae by the presence of a characteristic life cycle which includes the formation of a spherically dormant cyst. For the successful formation of the cyst both poly(hydroxybutyric acid) and alginate are essential. Alginate is an essential component of the cyst. This polysaccharide coating protects the cyst from desiccation and unfavourable conditions. The intine (inner coat) and exine (outer coat) layers of the cyst contain different types of alginate. The exine alginate is a polysaccharide with an unusually high content of consecutive guluronic acid residues (G-block), which results in a much higher stiffness. The intine alginate is rich in poly(mannuronic acid) residue. The existence of a large number of alginate-modifying enzymes (epimerases, lyases, and acetylase) in the differentiation process was reported for A. vinelandii (Holdal et al. 2000). Furthermore, the correlation between encystment and the release of epimerases indicates that epimerases are, additionally, cell- and cystassociated. Cysts have been reported to survive in dry soil for several years. Under favourable conditions, the alginate coating will swell and the cyst germinates (Skjak-Braek 1992; Nunez et al. 1999; Holdal et al. 2000; Young and Park 2007; Steigedal et al. 2008).
3.1.2
Alginate Capsule Formation as One of the Mechanisms of Nitrogenase Protection
The association of alginate production with cyst formation and its structural significance do not explain the abundant production of this polymer during vegetative growth of A. vinelandii. Indeed, under conditions permitting vegetative growth, alginate production plays different roles depending on the environmental conditions. A. vinelandii is a gram-negative bacterium, which fixes N2 non-symbiotically and aerobically. Among bacteria, aerobic nitrogen-fixing microbes are rare. They are mainly found in the family Azotobacteriaceae, which comprises the genera
Microbial Production of Alginates: Physiology and Process Aspects
157
Azotobacter, Azomonas, Beijerinchia, and Derxia. Pseudomonas methanitrificans, a methane-oxidizing organism also fixes N2 aerobically. Mycobacterium flavum and possibly one or two related species are also aerobic nitrogen fixers. The remainder of the aerobic bacteria which can fix N2 are facultative anaerobes which fix N2 only when they are grown anaerobically (Gottschalk 1988; Lichtl et al. 1997). The majority of nitrogen fixers belong to the group of strictly anaerobic microorganisms (Postgate 1971, 1974). The nitrogenase enzyme complex, which catalyses the reduction of N2 to ammonia, is highly sensitive to oxygen. However, members of the diazotrophic Azotobacter are able to grow under fully aerated conditions (Linkerhägner and Oelze 1995, 1997; Sabra et al. 1999, 2000). They do not fix nitrogen optimally at high dissolved oxygen concentrations (pO2) but tend to be microaerophilic in the sense that their N2 fixation is most effective at low pO2 though they are obligate aerobes. Obligate aerobes such as A. vinelandii are proposed to use two mechanisms for protecting the nitrogenase system against oxygen damage (Fig. 3) (Haddock and Jones 1977; Kuhla and Oelze 1988; Liu et al. 1995; Moshiri et al. 1995; Linkerhägner and Oelze 1997; Duyvis et al. 1998). The hypothesis of respiratory protection postulates that the unusually high activities of cellular oxygen consumption, characteristic for azotobacters, prevent the diffusion of O2 into the cells and thus to the nitrogenase. Whenever this steady state becomes disturbed by an increase in the pO2 or phosphate limitation, a conformational protection, i.e. the switch-off of nitrogenase activity,
Azotobacter cell +
N2
H2
Nitrogenase Mo Fe
NH3
Nitrogenase protection O2
+
PO4
Sucrose ATP
CO2
O2
+
PO4
CO2
Sucrose ATP
Active nitrogenase
Protected nitrogenase
Excess phosphate Active respiring cells
Phosphate limited Cell with inactive nitrogenase and ineffective oxidative phosphorylation
Fig. 3 Proposed mechanisms for the protection of nitrogenase against damage by oxygen in Azotobacter
158
W. Sabra and A.P. Zeng
protects nitrogenase proteins. Additional protection of the nitrogenase may also be provided by a decrease in the cellular surface area per cell volume as observed by Post et al. (1982). Previously, we observed that in a phosphate-limited continuous culture, the specific rate of oxygen consumption (qO2) of the cells increased as pO2 was elevated from 1 to 5% air saturation (Sabra et al. 2000). qO2 remained, however, essentially constant when pO2 was above 5%. These results cannot be interpreted in terms of the ‘respiratory protection’ concept. It was observed that both the molecular weight and the guluronic acid content of the alginate formed in these cultures monotonically increased with pO2. Transmission electron microscopy of negatively stained A. vinelandii showed that cells at lower and higher pO2 formed capsules with significant differences in the thickness and compactness of the polysaccharide (Fig. 4). These results revealed that under nitrogen-fixing conditions the bacterium tends to build alginate capsules with different composition according to the external pO2, implying that its production may contribute to a protection of nitrogenase against oxygen damage. Indeed, the results obtained in our laboratory (Sabra 1998; Sabra et al. 1999, 2000) and those obtained by Post et al. (1983) and Boiardi (1994) indicate that respiration, at least above a certain oxygen concentration, is not the prevailing mechanism for nitrogenase protection. The fact that nitrogen-fixing Azotobacter cells grown at their maximum specific oxygen consumption rate form compact capsules (Fig. 4), and that the intracellular oxygen level (CO2, int) in these nitrogen-fixing cells should not exceed the microaerobic- anaerobic range (Linkerhägner and Oelze 1997), indicates that the permeability of the cell wall to oxygen should decrease with increasing pO2 (CO2, ext) after reaching the maximum specific respiration rate (qO2 max). Under these conditions, the specific oxygen consumption rate (qO2) can be represented by qO2 = qO2 max = P (CO2,ext − CO2,int) = constant, where P is the permeability and is given by
pO2 = 2.5%
pO2 = 20%
PHB Cell membrane
Loose alginate slimy layer
Compact alginate capsule Bar = 0.5μm
Fig. 4 Thin-section electron micrograph of Azotobacter vinelandii cells grown at low (2.5% of air saturation) and high (20%) dissolved oxygen concentrations (pO2)
Microbial Production of Alginates: Physiology and Process Aspects
159
P = DO2 A/d. Here, DO2 represents the oxygen diffusion coefficient through the cell membrane (cm2s−1), A the specific cell surface area (cm2g−1), and d the cell wall thickness (cm). To achieve a constant qO2 value and to maintain a low intracellular oxygen concentration at an elevated level of pO2 in the medium, the organism could respond in three different ways according to the above equations: (1) decrease in the diffusion coefficient of oxygen (DO2) through a dense capsule, (2) decrease in the specific surface area (A); and (3) increase in the capsule thickness (d). All these will result in a decreased oxygen permeability. Post et al. (1982) reported a decrease in the specific surface area as a response to an increased oxygen concentration. The involvement of the first and third possibilities as protection mechanisms for the nitrogenase was suggested by Sabra et al. (2000, 2001). Studies on continuous culture of A. vinelandii growing diazotrophically with different oxygen concentrations indicated that alginate capsules are formed even in the presence of a high shear rate in a bioreactor. Moreover, the alginate capsule was more compact and dense at higher pO2 (Fig. 4). It was therefore postulated that the compact alginate layer formed around the cell acts also as a diffusion barrier for oxygen, which limits its transfer to the extremely oxygen sensitive nitrogenase enzyme. To finally confirm this, measurements of intracellular oxygen concentrations and quantitative knowledge of the sensitivity of nitrogenase towards oxygen are necessary. Figure 5 illustrates differences in the intracellular oxygen concentration as a function of pO2 for a normal aerobic microorganism with constant permeability and a nitrogen-fixing bacterium with varied permeability of the cell wall. In general, it is reasonable to state for vegetative cells of A. vinelandii that alginate has no single function, but rather provides the cell with a multitude of protective properties under various environmental conditions. Alginate can even act as a protective barrier against heavy metal toxicity, as an ion-exchange system with enhanced selectivity for Ca2+, or provide the bacterium with a hydrophilic, negatively charged coating which creates a barrier against attack and adverse environmental conditions (Fyfe and Govan 1983). Hammad (1998) revealed that alginate-encapsulated Azotobacter chrococcum is better protected against the depressive effect of phages than are non-immobilized cells. The presence of eight different types of epimerase together with lyases and acetylase represent excellent tools for the cell to manufacture polymers of different and unique properties to cope with the external environment.
3.2
Pseudomonads
3.2.1
Role of Alginate in Biofilm Formation
Alginate is a key component of biofilm of P. aeruginosa and certain other pseudomonads (see “Role of Alginate in Bacterial Biofilms” by Halverson, this volume). Alginate biosynthesis and biofilm formation can be triggered by a number of
160
W. Sabra and A.P. Zeng Saturation range Permeability (P) qO2 Aerobic bacterium
r la
Permeability
llu
ce
ra
2
t in
CO
CO2
Permeability (P)
qO2
qO2 Aerobic N2 fixing bacterium
CO2 intracellular
0
20
40
60
80
pO2 (CO2 extracellular)
Fig. 5 Differences in the intracellular O2 concentration as function of pO2 in culture for a normal aerobic bacterium with constant permeability and a nitrogen-fixing bacterium with varied permeability of the cell wall
environmental factors, including high osmolarity, ethanol, nitrogen, or phosphate limitation (Govan and Deretic 1996; Wagner and Iglewski 2008). It is recognized that oxygen plays an important role in the formation of alginate and mucoidy appearance in this bacterium (Sato et al. 1984; Krieg et al. 1986; Bayer et al. 1990; Leitäo and Sa-Correia 1997; Xu et al. 1998). Oxygen-dependent upregulation of transcription of alginate genes was normally found in highly mucoid strains of P. aeruginosa (Leitäo and Sa-Correia 1997). In an experiment with a mixture of mucoid and non-mucoid strains (1:1), Krieg et al. (1986) found that aeration favourably selects the growth of only mucoid phenotypes. Non-mucoid strains were reported to be sensitive to oxygen. In this connection it should also be mentioned that not all cystic fibrosis isolates that appear non-mucoid are truly non-mucoid, but, instead, may belong to the stress-conditio -dependent category of mucoid mutants. Recently, it was shown that in a non-mucoid PAO1 strain of P. aeruginosa, alginate formation can be triggered in chemostat culture under oxidative stress conditions (Sabra et al. 2002; Kim et al. 2003). Interestingly, and despite the lack of the oxygensensitive nitrogenase in P. aeruginosa, this bacterium exhibits better growth in the microaerophilic range as A. vinelandii, (Sabra et al. 2000, 2001, 2002).
Microbial Production of Alginates: Physiology and Process Aspects
161
The preference of P. aeruginosa for microaerophilic growth is in accordance with findings from genomic analysis for this strain (Croft et al., 2000). Analysis of the genomic DNA sequences of P. aeruginosa PAO1 showed that this strain possesses a 15-kb cluster of genes encoding a microoxic respiration system, similar to that used by nitrogen-fixing bacteria, including genes encoding homologues of nitrogen fixation/microoxic regulatory cascade proteins (fixL, fixJ, and fixK/anr). These genes enable the preferred microaerobic growth of nitrogen-fixing bacteria such as A. vinelandii. For P. aeruginosa, they might help in its unusual survival under the conditions in the biofilm (Boyd and Chakrabarty 1995). We further investigated the mechanisms by which P. aeruginosa protects itself against oxidative stress (Sabra et al. 2002; Kim et al 2003). The P. aeruginosa strain PAO1 can create microaerophilic growth conditions by at least two mechanisms: (1) reduction of the transfer rate of oxygen and (2) formation of a polysaccharide capsule on the cell surface (Sabra et al. 2002; Kim et al. 2003; Sabra et al. 2003). It is postulated that the blockage of oxygen transfer may play an important role in the defence of this pathogen against reactive oxygen intermediates. Figure 6 shows the difference between Pseudomonas cells grown under nearly anaerobic conditions, in flasks, and in a pO2-controlled chemostat culture. The difference in capsule morphology of P. aeruginosa from that of A. vinelandii may reflect the variations in capsule structures, in particular the occurrence of a G-G-block in the latter, and hence the production of a compact alginate layer around the nitrogen-fixing bacterium (Figs. 4, 6)
3.2.2
Alginate and Mucoidy Prevalence in Clinical P. aeruginosa Strains
Over the years, a prime target for antimicrobial therapy has been P. aeruginosa. However, despite constant improvements in antibiotic research, P. aeruginosa remains impossible to eradicate. The association of mucoid strains of P. aeruginosa with chronic pulmonary infection in cystic fibrosis patients is well recognized (Pritt et al. 2007). Evidence for their emergence in vivo was first proven by Doggett et al. (1966). However, in contrast to their emergence in vivo, when they are cultured in vitro, mucoid strains of P. aeruginosa tend to revert to the non-mucoid form. The particua
b
0.8 μm Cell membrane
Polysaccharide capsule
Fig. 6 Transmission electron microscopy of ultrathin sections of negatively stained Pseudomonas aeruginosa cells grown under different conditions: a cells from shake flask culture (pO2 = 0); b cells grown under controlled oxygen pressure (pO2 = 10% of air saturation)
162
W. Sabra and A.P. Zeng
lar cause for the conversion to mucoidy is not fully understood in P. aeruginosa. In the course of phage typing, Martin (1973) observed mucoid growth of P. aeruginosa immediately around areas of phage lysis and obtained mucoid variants after subculturing from such areas. Prolonged antibiotic therapy has been proven as a causative factor in the emergence of mucoidy in patients with cystic fibrosis (Govan and Fyfe 1978; Alkawash et al. 2006; Wood et al. 2006; Prett et al. 2007; Goh et al. 2008). Indeed, isolation of alginate-producing mutants of P. fluorescens, P. putida, and P. mendocina was done through treatment with antibiotics. Moreover, in vivo, aged biofilms with a high concentration of alginate materials are significantly less susceptible to antibiotic therapy (Anwar et al. 1992; Alkawash et al. 2006). Oxidative burst in neutrophils in contact with P. aeruginosa in the lung and the release of oxidative radicals was reported to be another reason for the emergence of mucoidy (Mathee et al. 1999; Sabra et al. 2002, Kim et al. 2003).
3.2.3
Other Pseudomonads
For non-pathogenic alginate-producing pseudomonads no natural ecological niche was reported. This suggests that in the majority of natural habitats, alginate biosynthesis provides no advantage to the organism. Non-mucoid pseudomonads are routinely isolated from natural environments. Since EPS are known to chelate heavy metals, it was also reported that alginate formation was induced, in plant-parasitic pseudomonads, by treatment with a bacteriocidal spray containing copper. Thus, the secretion of alginate may contribute to some plant-bacterial diseases. Although alginate is an important virulence factor in the infection of plants by some phytopathogenic pseudomonads, the precise purpose of this polysaccharide remains unclear. Especially for some pseudomonads that produce more than one polysaccharide. For example, P. syringae pv. glycinea produces levan and acetylated alginates in vitro when grown on sucrose or glucose as the carbon source, respectively, yet alginate is the only polysaccharide isolated from the infected soybean leaves (Osman et al. 1986; Laue et al. 2006; Schenk et al. 2008). Sodium chloride and ethanol were shown to significantly increase alginate production in a variety of fluorescent pseudomonads, suggesting that osmolarity and dehydration may be general signals for the production of this polysaccharide (Kidambi et al. 1995; Chang et al. 2007).
4 4.1
Optimal Conditions for Microbial Alginate Production Fermentation Conditions for Optimal Alginate Production in A. vinelandii
Attempts to increase alginate productivity through medium formulation were the focus of many studies (Horan et al. 1981; Okabe et al. 1981; Lebrun et al. 1994; Savalgi and Savalgi 1992; Pena et al. 1997; Clementi et al. 1999; Seanez et al. 2001;
Microbial Production of Alginates: Physiology and Process Aspects
163
Saude et al. 2002; Saude and Junter 2002: Trujillo-Roldán et al. 2003; Díaz-Barrera et al. 2007; Galindo et al. 2007; Pena et al. 2008). The addition of fixed nitrogenous compounds to the culture medium was reported controversially in the literature. Brivonese and Sutherland (1989) concluded that varying the source of peptone used in the medium could alter alginate yield by up to 30%, suggesting a more specific role for nitrogenous nutrients. Furthermore, growth in the presence of ammonium inhibits alginate formation in A. vinelandii (Brivonese and Sutherland 1989; Sabra 1998). To understand the precise role of nitrogenous compounds in alginate production, pO2 must be accurately controlled. However, most of the work in this respect was done in uncontrolled shake flask cultures. This may explain the contradictory results obtained with A. vinelandii. The crucial effect of controlling pO2 explains also the controversially reported results with respect to phosphate (Deavin et al. 1977; Jarman et al. 1978; Chen et al. 1985; Sabra et al. 1999). On the one hand, Brivonese and Sutherland (1989) reported that phosphate might simply act to buffer the medium and used a phosphaterich medium (7.5g K2HPO4). On the other hand, observed a drop in alginate yield in a medium with excess phosphate was observed and for their continuous culture studies phosphate limitation gave the maximum alginate production (Sabra et al. 1999). Moreover a correlation between phosphate limitation, respiratory requirement of the cell in term of the respiratory quotient (RQ), and alginate production by diazotrophically grown cells of A. vinelandii was demonstrated by Sabra et al. (1998, 1999). They showed that in phosphate-limited batch or continuous culture, the highest specific alginate production was accompanied by a RQ value around 0.8 at an intermediate value of pO2 (2–5% air saturation). This RQ value was, indeed, near the theoretical optimal RQ value (0.8) calculated on the basis of the stoichiometry of bioreactions leading to alginate formation from sucrose. This optimal RQ appeared to be a useful parameter for the control of alginate production by A. vinelandii. It is generally recognized that the control of oxygen supply is of critical importance for alginate production by A. vinelandii especially when it is grown diazotrophically. We showed previously that, under controlled microaerophilic conditions, a very narrow range of pO2 exists for both optimal alginate and biomass production under diazotrophic conditions (Sabra et al. 1999). This extreme oxygen sensitivity renders the studies on cell physiology with respect to the effects of different nutritional conditions difficult and sometimes controversial (see above). Indeed, it was shown that under phosphate-limited conditions, results obtained under controlled microaerophilic conditions differed significantly from those of flask cultures with the same phosphate concentrations (Sabra et al. 1999). The effect of increasing the viscosity of the culture broth due to the polymer production and cell growth on the effective pO2, on the one hand, and the decrease in pH due to alginate secretion, on the other hand, make such heterogeneous environments in uncontrolled culture unsuitable for investigating the physiology of such a process. Several studies on Azotobacter showed that for optimal alginate formation, pO2 must be accurately controlled in a microaerophilic range. Oxygen limitation in this bacterium leads to increased accumulation of poly(b-hydroxybutyrate) (Horan et al. 1983; Parente et al. 1998; Sabra et al. 2000; Noguez et al. 2008). The narrow microaerophilic range of
164
W. Sabra and A.P. Zeng
alginate production by a diazotrophically growing culture of Azotobacter remains a challenging issue in large-scale production. Recently, it was concluded that alginate production does not increase linearly with increasing pO2, owing to the fact that the dense alginate capsule around the cell (at high pO2) exhibits not only a mass-transfer barrier against oxygen but also against other medium constituents and hence lowers the biomass and alginate productivity. In contrast, the soft alginate slimy layer at intermediate pO2 exhibits lower mass-transfer resistance to other nutrients and thus permits the continuous production of alginate (Sabra et al. 2000). The effect of other growth limitations on alginate production has also been thoroughly investigated. Iron and/or molybdate limitation gave the highest specific rates for alginate production (Deavin et al. 1977; Annison and Couperwhite 1986b; Ferrala et al. 1986). Ca2+ limitation, on the other hand, did not enhance alginate and biomass formation in diazotrophically grown cells (Annison and Couperwhite 1986b; Obika et al. 1993). This may reflect the importance of epimerase activity, which is known to be dependant on the Ca2+ level, for the survival of these cells. The chelation of Ca+2 through the addition of EDTA to the fermentation medium resulted in the production of alginate with a preponderance of mannuronic acid residues (Couperwhite and McCallum 1974; Annison and Couperwhite 1986a). Surprisingly, under sucrose limitation in a fed-batch culture, conditions where the cell would be expected to make the most efficient use of its available carbon and energy substrate, alginate was produced at similar rates as for other limitations (Sabra 1998; Sabra et al. 2001). This was also observed in carbon-limited chemostat culture by Deavin et al. (1977). The formation of alginate under carbon limitation may indicate how important alginate biosynthesis is for the survival of this diazotrophically grown bacterium in controlled pO2 culture. This means that the protection of nitrogenase through the formation of a compact capsule around the cell under carbon limitation is the predominating mechanism for respiratory protection, so wasting the limited carbon source is not advantageous for cell survival. In agreement with this assumption, the guluronic acid content and the relative molecular weight of alginate produced after the onset of carbon limitation in a fedbatch culture increased (Sabra 1998). Although the carbon-limited culture produced alginate of high quality (higher molecular weight and higher guluronic acid content), a lower yield was obtained in comparison with that of a normal batch process (3 and 4.9gl−1 for fed-batch and batch culture, respectively) (Sabra 1998).
4.2
Alginate Production by P. aeruginosa
The fluorescent pseudomonads have the potential to produce many different types of EPS constituents, including, for example, levan, marginalan, cellulose, and alginate, in addition to several uncharacterized polymers (Fialho et al. 1990; Fett et al. 1995; Kachlany et al. 2001; Jackson et al. 2004; Laue et al. 2006; Priester et al. 2006). Despite the nearly universal ability of Pseudomonas species to produce alginate, P. aeruginosa was the one used in most studies (Qiu et al. 2008; Stehling et al. 2008), because of its association with virulence in cystic fibrosis patients. Indeed, identifying environmental factors influencing the expression of alginate by the
Microbial Production of Alginates: Physiology and Process Aspects
165
opportunistic pathogen may help to understand the process of pathogenesis. Moreover, it was recently concluded that mucoid and non-mucoid phenotypes can be expressed according to the host’s tissues or environment (Stehling et al. 2008). It has been well recognized that oxygen plays an important role in the formation of alginate and mucoidy appearance in P. aeruginosa (Sato et al. 1984; Krieg et al. 1986; Leitäo and Sä-Correia 1997; Xu et al. 1998; Sabra et al. 2002; Kim et al. 2003). Oxygen-dependent upregulation of transcription of alginate genes was normally found in highly mucoid strains of P. aeruginosa (Leitäo and Sä-Correia 1997). Moreover, in an experiment with a mixture of mucoid and nonmucoid strains (1:1), (Krieg et al. 1986) found that aeration favourably selected the growth of mucoid phenotypes. In fact, we reported recently that oxidative stress may cause the conversion to mucoidy (Sabra et al. 2002). Generally, alginate production in vitro is increased upon nutrient (carbon, phosphate, nitrogen, iron) deprivation or exposure to increased sodium chloride, ethanol, desiccation, and oxygen (Ma et al. 1998; Kim et al. 2003; Chang et al. 2007).
4.3
Kinetics of Alginate Production
Alginate production by A. vinelandii in phosphate-limited batch cultures was found to parallel cell growth and to cease when growth stopped (Sabra et al. 1999). This was in agreement with the results of Deavin et al. (1977). The first attempt to describe the kinetics of alginate production was made by Klimek and Ollis (1980). They used the experimental data from Deavin et al. (1977) in batch cultures to describe the rate of alginate production by the well-known Luedeking and Piret kinetics (Luedeking and Piret 1959), and a growth-associated form allowed this fermentation processes to be described well. However, in continuous culture, they found that the same strains showed a specific alginate production rate that was relatively independent of growth rate (Deavin et al. 1977; Jarman et al. 1978). These controversial results for batch and continuous cultures were later explained by the fact that alginate synthesis might have been inhibited in the late stage of the batch culture by oxygen limitation that did not occur in continuous culture, owing to pO2 control (Margaritis and Pace 1985). The contribution of the non-growing-cell metabolism to alginate production was observed by Clementi et al. (1995) using A. vinelandii DSMZ 576, and alginate production continued after the organisms entered the stationary phase. They also confirmed that alginate production was only partially associated with growth, with a 75–80% contribution of nongrowth-associated production. Recently, we showed that for A. vinelandii DSMZ 93–541b at a fixed growth rate the specific alginate production rate (qalg) was dependent on the pO2 applied (Fig. 7). Furthermore, a general trend was observed, namely the qalg peaked at intermediate pO2 (2–5% air saturation). It may, therefore, be concluded that the kinetics of alginate production depends not only on the growth rate (Brivonese and Sutherland 1989; Sabra et al. 2000) but also on other culture conditions. Because of the association of P. aeruginosa with pathogenesis, the results of very few kinetic studies have been published on this organism (Krieg et al. 1986). Moreover, most of the work reported in the literature concerning alginate production, if any, was done in petri plates or uncontrolled flask cultures (Remminghorst
166
W. Sabra and A.P. Zeng 0.26 0.24 0.22
(D = 0.08) (D = 0.22)
(D = 0.15) (D =0.26)
0.20
qalg (g/g∗h)
0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 1
2.5
5
10
pO2 value (%)
Fig. 7 Dependence of the specific alginate production rate (qalg; grams of alginate per gram of biomass per hour) on pO2 at the same growth rate of A. vinelandii grown in phosphate-limited chemostat culture
and Rehm 2006; Qiu et al. 2008; Stehling et al. 2008). Indeed, there is a lack of kinetics data, especially for those clinical isolates of Pseudomonas. The instability of alginate production in vitro by most Pseudomonas strains represents a major burden for data evaluation or kinetic analysis (Schurr et al. 1994). It was earlier reported that a mucoid strain of P. aeruginosa converted to a non-mucoid phenotype upon prolonged cultivation in continuous culture (Krieg et al. 1986). Recently, Sabra and Hassan (2008) examined the behaviour of alginate production kinetics by a stable mucoid sputum isolate from a cystic fibrosis patient (FRD1). With use of a statistical method, through factorial experimental design and response surface methodology (Yuan et al. 2008), it was shown that alginate production by the FRD1 strain exhibits a typical secondary metabolite production pattern with the highest production rate at the lowest growth rate (Fig. 8). The diminished growth rate that favours alginate production by the clinical strain of P. aeruginosa may simulate the condition in the biofilm where the bacteria are most persistent.
4.4
Non-Newtonian Alginate Fluids and the Role of Fermentor Hydrodynamics
In aerobic submerged exocellular microbial polysaccharides, fermentations such as those of xanthan, dextran and alginate, oxygen supply to the media, and mass transfer of oxygen to growing microbial cells are still a major technical problem affecting
Specific alginate formation rate (g/g∗h)
Microbial Production of Alginates: Physiology and Process Aspects
167
0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.20
0.21
0.22
0.23
0.24
0.25
0.26
0.27
0.28
Specific growth rate (h–1)
Fig. 8 Specific alginate production rate (qalg) as a function of the growth rate of P. aeruginosa FRD1 grown in batch cultures designed according to response surface methodology
microbial productivity, since the solution becomes highly viscous and non-Newtonian during fermentation owing to the presence of polysaccharides in the culture medium. The relationship between energy input and oxygen transfer rate in a bioreactor with pseudoplastic behaviour is not easy to establish (Dussap and Gros 1985; Pena et al. 2002; Zhang and Zhong 2004; Garcia-Ochoa and Gomez 2005). In the case of alginate production by A. vinelandii, the low oxygen solubility in fermentation media coupled with the high oxygen consumption rate make the task of ensuring there is sufficient oxygen supply (aeration) difficult. Changing the rheological properties of the fluid will obviously affect the flow pattern created by the impeller. Around the impeller the fluid is turbulent, and becomes laminar or stagnant when the shear stresses are below the yield stress of the polymer. The impeller creates a cavern in which the fluid moves relatively fast and where the flow is turbulent. However, in the bulk of the fluid, where the shear stresses are below the yield stresses, the fluid does not move at all (McNeil and Harvey 1993; Mancini et al. 1996; Amici et al. 2008). Preliminary experiments in laboratory bioreactors used for the production of microbial alginate were typically restricted to aerated, mechanically agitated vessels (Deavin et al. 1977; Sabra et al. 2000; Trujillo-Roldan et al. 2001). This gave the turbulence necessary for small bubble formation and mixing and hence good mass transfer. However, for the oxygen-sensitive A. vinelandii, the necessity of accurately controlling the pO2 in large-scale bioreactors remains a challenging problem. The pseudoplastic behaviour of alginate solution results in an exponential increase in viscosity as the fluid slows down after it leaves the impeller and hence the reduction in the pO2. On the other hand, the relatively low viscosity in the
168
W. Sabra and A.P. Zeng
region near the impeller tends to promote channelling of gas up the centre of the vessel, creating a highly oxygenated zone. To our knowledge, experiments with other types of reactor, especially bubble column and air lift reactors, have not been done for alginate and are worth testing. In fact, for xanthan fermentation, a bubble column turned out to be more efficient than a stirred tank bioreactor on a large scale (Deckwer 1992; Suh et al. 1992). There are very few reports covering aspects of scale-up of the alginate production process (Trujillo-Roldan et al. 2001; Reyes et al. 2003). Understanding the operational parameters involved in scaling-up of alginate production is important. The viscosity and the molecular weight of the polymer produced by A. vinelandii in shake flasks were reported to be greater than those of the polymer produced in a pO2-controlled bioreactor at the same polymer concentration (Pena et al. 2000). Recently Pena et al. (2008) succeed in achieving a high molecular weight alginate polymer in a stirred bioreactor similar to that obtained in flasks by simulating the evolution of actual power input occurring in shake flasks.
5
Conclusion
Commercial opportunities seem to exist for bacterial alginates, especially in the biomedical and pharmaceutical fields. The high price of alginates applied in these fields may open up a new market for polymers with a high degree of purity and defined chemical compositions. Firm knowledge of the relationship between alginate structure and function is needed to produce a polymer with the required structure for specific applications. In this regard and in view of possible industrial production of alginate, there are several aspects which need further investigation. For A. vinelandii, the problem of oxygen sensitivity represents a challenging task for large-scale production in a bioreactor. The effect of reactor design on the production and the quality of alginate is still not clear. An alternative physiological solution is to decrease the extreme oxygen sensitivity of this bacterium. This could enhance the feasibility of commercialized production of alginate from A. vineladii. It can be stated that the microbial production of high-quality alginate is technically and ecologically more favourable than the algal route. In this respect, understanding the mechanisms of epimerization, acetylation, and degradation, as well as their regulation in both A. vinelandii and P. aeruginosa is crucial for obtaining a highquality product for a potentially new market.
References Alkawash MA, Soothill JS, Schiller NL (2006) Alginate lyase enhances antibiotic killing of mucoid Pseudomonas aeruginosa in biofilms. APMIS 114(2):131–138 Amici E, Tetradis-Meris G, Pulido de Torrres C, Jousse F (2008) Alginate gelation in microfluidic channels. Food Hydrocolloids 22:97–104
Microbial Production of Alginates: Physiology and Process Aspects
169
Annison G, Couperwhite I (1986a) Influence of calcium on alginate production and composition in continuous cultures of Azotobacter vinelandii. Appl Microbiol Biotechnol 25:55–61 Annison G, Couperwhite I (1986b) Effect of limiting substrate concentration, growth rate and aeration on alginate composition and production by Azotobacter vinelandii in continuous culture. Food Hydrocolloids 1(2):101–111 Anwar H, Strap JL, Costerton JW (1992) Establishment of aging biofilms: possible mechanisms of bacterial resistance to antimicrobial therapy. Antimicrob Agent Chemother 36:1347–1351 Bayer AS, Eftekhar F, Tu J, Nast CC, Speert DP (1990) Oxygen-dependent up- regulation of mucoid exopolysaccharide (alginate) production in Pseudomonas aeruginosa. Infect Immun 58(5):1344–1349 Boiardi JL (1994) Metabolic cost of nitrogen incorporation by N2-fixing Azotobacter vinelandii is affected by the culture pH. Biotechnol Lett 16(11):1195–1198 Boyd A, Chakrabarty AM (1995) Pseudomonas aeruginosa biofilms: Role of the alginate exopolysaccharide. J Ind Microbiol 15:162–168 Brivonese A, Sutherland WI (1989) Polymer production by a mucoid strain of Azotobacter vinelandii in batch culture. Appl Microbiol Biotechnol 30:97–102 Chang WS, van de Mortel M, Nielsen L, Nino de Guzman G, Li X, Halverson LJ (2007) Alginate production by Pseudomonas putida creates a hydrated microenvironment and contributes to biofilm architecture and stress tolerance under water-limiting conditions. J Bacteriol 89(22):8290–8299 Chen WP, Chen JY, Chang SC, Su CL (1985) Bacterial alginate produced by a mutant of Azotobacter vinelandii. Appl Environ Microbiol 49(3):543–546 Clementi F, Fantozzii P, Mancini F, Moresi M (1995) Optimal conditions for alginate production by Azotobacter vinelandii. Enz Microbiol Technol 17(11):983–988 Clementi F, Crudele MA, Parente E, Mancini M, Moresi M (1999) Production and characterisation of alginate from Azotobacter vinelandii. J Sci Food Agric 79:602–610 Couperwhite I, McCallum MF (1974) The influence of EDTA on the composition of alginate synthesized by Azotobacter vinelandii. Arch Microbiol 97:73–80 Croft L, Beatson SA, Whitchurch CB, Huang B, Blakeley RL, Mattick JS (2000) An interactive web-based Pseudomonas aeruginosa genome database: Discovery of new genes, pathways and structures. Microbiology 146:2351–2364 Deavin L, Jarman TR, Lawson CJ, Richelato RC, Slocombe S (1977) The production of alginic acid by Azotobacter vinelandii in batch and continuous culture. In: Sanford PA, Laskin A (eds) Extracellular microbial polysaccharides. American Chemical Society, Washington, pp 14–26 Deckwer W-D (1992) Bubble column reactors. Wiley, Chichester Díaz-Barrera A, Peña C, Galindo E (2007) The oxygen transfer rate influences the molecular mass of the alginate produced by Azotobacter vinelandii. Appl Microbiol Biotechnol 76(4): 903–910 Doggett RG, Harrison GM, Stillwell RM, Wallis ES (1966) An atypical Pseudomonas aeruginosa associated with cystic fibrosis of the pancreas. J Pediatr 68:215–221 Dussap C-G, Gros J-B (1985) Power input, KLa values in pneumatically agitated fermentors with extracellular microbial polysaccharides. In: Biotechnologie 85, Europe, pp 691–692 Duyvis M, Wassink H, Haaker H (1998) Nitrogenase of Azotobacter vinelandii: kinetic analysis of the Fe protein redox cycle. Biochemistry 37:17345–17354 Ferrala NF, Westervelt P, Mabbott GA, Fekete FA (1986) Relation between extracellular polysaccharide production and medium iron concentration in nitrogen fixing Azotobacter chrococcum B-8. Abstr Annu Meet Am Soc Microbiol 86:Meeting K-143 Fett WF, Wells JM, Cescutti P, Wijey C (1995) Identification of exopolysaccharides produced by fluorescent pseudomonads associated with commercial mushroom (Agaricus bisporus) production. Appl Environ Microbiol 61:513–517 Fialh AM, Zielinski NA, Fett WF, Chakrabarty AM, Berry A (1990) Distribution of alginate gene sequences in the Pseudomonas rRNA homology group I-Azomonas-Azotobacter lineage of superfamily B procaryotes. Appl Environ Microbiol 56:436–443
170
W. Sabra and A.P. Zeng
Fyfe JAM, Govan JRW (1983) Synthesis, regulation and biological function of bacterial alginate. Prog Ind Microbiol 18:45–83 Galindo E, Peña C, Núñez C, Segura D, Espín G (2007) Molecular and bioengineering strategies to improve alginate and polydydroxyalkanoate production by Azotobacter vinelandii. Microb Cell Fact 6:7 Garcia-Ochoa F, Gomez E (2005) Prediction of gas-liquid mass transfer coefficient in sparged stirred tank bioreactors. Biotechnol Bioeng 92(6):761–772 Goh CH, Heng PW, Huang EP, Li BK, Chan LW (2008) Interactions of antimicrobial compounds with cross-linking agents of alginate dressings. J Antimicrob Chemother 62(1):105–108 Gorin JPA, Spencer TJF (1966) Exocellular alginic acid from Azotobacter vinelandii. Can J Chem 44:993–998 Gottschalk G (1988) Fixation of molecular nitrogen. In: Gottschalk G (ed) Bacterial metabolism, vol 10. Springer, Berlin, pp 318–326 Govan JR, Deretic V (1996) Microbial pathogensis in cystic fibrosis: Mucoid Pseudomonas aeruginosa and Burkholderia cepacia. Microbiol Rev 60(3):539–574 Govan JR, Fyfe MJA (1978) Mucoid Pseudomonas aerugionosa and cystic fibrosis: resistance of the mucoid form to carbencillin, flucloxacillin and tobramycin and the isolation of mucoid variants in vitro. J Antimicrob Chemother 4:233–240 Haddock BA, Jones CW (1977) Bacterial respiration. Bacteriol Rev 41(1):47–99 Hammad AMM (1998) Evaluation of alginate encapsulated Azotobacter chroococcum as a phageresistant and an effective inoculum. J Basic Microbiol 1:9–16 Holdal HK, Svanem IG, Gimmestad M, Valla S (2000) Mannuronan C-5 epimerases and cellular differentiation of Azotobacter vinelandii. Environ Microbiol 2(1):27–38 Horan NJ, Jarman TR, Dawes EA (1981) Effects of carbon source and inorganic phosphate concentration on the production of alginic acid by a mutant of Azotobacter vinelandii and on the enzyme involved in its biosynthesis. J Gen Microbiol 127:185–191 Horan NJ, Jarman TR, Dawes EA (1983) Studies on some enzymes of alginic acid biosynthesis in Azotobacter vinelandii grown in continuous culture. J Gen Microbiol 129:2985–2990 Jackson KD, Starkey M, Kremer S, Parsek MR, Wozniak DJ (2004) Identification of psl, a locus encoding a potential exopolysaccharide that is essential for Pseudomonas aeruginosa PAO1 biofilm formation. J Bacteriol 186:4466–4475 Jarman TR, Deavin I, Slocombe S, Righelato RC (1978) Investigation of the effect of environmental conditions on the rate of exopolysaccharides synthesis in Azotobacter vinelandii. J Gen Microbiol 107:59–64 Kachlany SC, Levery SB, Kim JS, Reuhs BL, Lion LW, Ghiorse WC (2001) Structure and carbohydrate analysis of the exopolysaccharide capsule of Pseudomonas putida G7. Environ Microbiol 3:774–784 Kidambi PS, Sundin GW, Palmer AD, Chakrabarty MA, Bender LC (1995) Copper as a signal for alginate synthesis in Pseudomonas syringae pv. syringae. Appl Environ Microbiol 61(6): 2172–2179 Kim EJ, Sabra W, Zeng AP (2003) Iron deficiency leads to inhibition of oxygen transfer and enhanced formation of virulence factors in cultures of Pseudomonas aeruginosa PAO1. Microbiology 149(9):2627–2634 Klimek J, Ollis DF (1980) Extracellular microbial polysaccharides: kinetics of Pseudomonas sp, Azotobacter vinelandii and Aureobasidium pullulans batch fermentations. Biotechnol Bioeng 22:2321–2342 Krieg DP, Bass JA, Mattingly SJ (1986) Aeration selects for mucoid phenotype of Pseudomonas aeruginosa. J Clin Microbiol 24(6):986–990 Kuhla J, Oelze J (1988) Dependence of nitrogenase switch-off upon oxygen stress on the nitrogenase activity in Azotobacter vinelandii. J Bacteriol 170(11):5325–5329 Laue H, Schenk A, Li H, Lambertsen L, Neu TR, Molin S, Ullrich MS (2006) Contribution of alginate and levan production to biofilm formation by Pseudomonas syringae. Microbiology 152(10):2909–2918 Lebrun L, Junter G-A, Jouenne T, Mignot L (1994) Exopolysaccharide production by free and immobilized microbial cultures. Enzyme Microb Technol 16:1048–1054
Microbial Production of Alginates: Physiology and Process Aspects
171
Leitäo JH, Sa-Correia I (1997) Oxygen dependant upregulation of transcription of alginate genes algA, algC and algD in Pseudomonas aeruginosa. Res Microbiol 148:37–43 Lichtl RJ, Bazin MO, Hall D (1997) The biotechnology of hydrogen production by Nostoc flaelliforme grown under chemostat conditions. Appl Microbiol Biotechnol 47:701–707 Linker A, Jones RS (1966) A polysaccharide resembling alginic acid from a Pseudomonas microorganism. Nature 204:187–188 Linkerhägner K, Oelze J (1995) Cellular ATP level and nitrogenase switchoff upon oxygen stress in chemostat cultures of Azotobacter vinelandii. J Bacteriol 177(18):5289–5293 Linkerhägner K, Oelze J (1997) Nitrogenase activity and regeneration of the cellular ATP pool in Azotobacter vinelandii adapted to different oxygen concentrations. J Bacteriol 179(4):1362–1367 Liu JK, Lee FT, Lin CS, Yao XT, Davenport JW, Wong TY (1995) Alternative function of the electron transport system in Azotobacter vinelandii: Removal of excess reductant by the cytochrome d pathway. Appl Environ Microbiol 61(11):3998–4003 Luedeking R, Piret EL (1959) A kinetic study of lactic acid fermentation. Batch process at controlled pH. J Biochem Microbial Technol Eng 1:393–431 Ma S, Selvaraj U, Ohman DE, Quarless R, Hassett DJ, Wazniak DJ (1998) Phosphorylationindependent activity of the response regulators AlgB and AlgR in promoting alginate biosynthesis in Mucoid Pseudomonas aeruginosa. J Bacteriol 180(4):956–968 Mancini M., Moresi M, Sappino F (1996). Rheological behaviour of aqueous dispersions of algal sodium alginates. J Food Eng 28:238–295 Margaritis A, Pace GW (1985) Microbial polysaccharides. In: Blanch HW, Drew S, Wang DIC (eds) Comprehensive biotechnology, vol 3. Pergmann, Oxford, pp 1005–1044 Martin DR (1973) Mucoid variation in Pseudomonas aeruginosa induced by the action of phage. J Med Microbiol 6:111–118 Mathee K, Ciofu O, Sternberg C, Lindum PW, Campbell JIA, Jensen P, Johnsen A, Givskov M, Ohman DE, Molin S, Hoiby N, Kharazmi A (1999) Mucoid conversion of Pseudomonas aeruginosa by hydrogen peroxide: A mechanism for virulence activation in the cystic fibrosis lung. Microbiology 145:1349–1357 McNeil B, Harvey LM (1993) Viscous fermentation products. Crit Rev Biotechnol 13(4):275–304 Moshiri F, Crouse BR, Johnson MK, Maier RJ (1995) The nitrogenase protective FeSII protein of Azotobacter vinelandii: Over expression, characterisation and crystallisation. Biochemistry 34:12973–12982 Noguez R, Segura D, Moreno S, Hernandez A, Juarez K, Espín G (2008) Enzyme I, NPr and IIA(Ntr) are involved in regulation of the poly-beta-hydroxybutyrate biosynthetic genes in Azotobacter vinelandii. J Mol Microbiol Biotechnol 15:244–254 Nunez C, Moreno S, Soberon-Chavez G, Espin G (1999) The Azotobacter vinelandii response regulator AlgR is essential for cyst formation. J Bacteriol 181(1):141–148 Obika H, Sakakibara J, Kobayashi Y (1993) Direct control of the constituent ratio in a wide range in alginate produced by Azotobacter vinelandii. Biosci Biotechnol Biochem 57(2):332–333 Okabe E, Nakayima M, Murooka H, Nisizawa K (1981) Investigation of carbon and phosphorous sources in cultural media of a selected strain of alginate producing Azotobacter vinelandii. J Ferment Technol 59(1):1–7 Osman SF, Fett WF, Fishman ML (1986) Exopolysaccharides of the phytopathogen Pseudomonas syringae pv. glycinea. J Bacteriol 166:66–71 Parente E, Crudele MA, Aquino M, Clementi F (1998) Alginate production by Azotobacter vinelandii DSM576 in batch fermentation. J Ind Microbiol Biotechnol 20:171–176 Pena C, Campos N, Galindo E (1997) Evolution of alginate molecular weight distributions, broth viscosity and morphology of Azotobacter vinelandii cultured in shake flasks. Appl Microbiol Biotechnol 48:510–515 Pena C, Trujillo-Roldan MA, Galindo E (2000) Influence of dissolved oxygen tension and agitation speed on alginate production and its molecular weight in cultures of Azotobacter vinelandii. Enzyme Microb Technol 27:390–398 Peña C, Galindo E, Díaz M (2002) Effectiveness factor in biological external convection: study in high viscosity systems. J Biotechnol 95(1):1–12
172
W. Sabra and A.P. Zeng
Pena C, Milla M, Galindo E (2008). Production of alginate by Azotobacter vinelandii in a stirred fermentor simulating the evolution of power input observed in shake flasks. Process Biochem 43:775–778 Post E, Golecki RJ, Oelze J (1982) Morphological and ultrastructural variations in Azotobacter vinelandii growing in oxygen-controlled continuous culture. Arch Microbiol 133:75–82 Post E, Kleiner D, Oelze J (1983) Whole cell respiration and nitrogenase activities in Azotobacter vinelandii growing in oxygen controlled continuous culture. Arch Microbiol 134:68–72 Postgate JR (1971) Nitrogen fixation by free-living microbes. In: Postgate JR (ed) The chemistry and biochemistry of nitrogen fixation, vol 5. Plenum, London, pp 161–187 Postgate JR (1974) Evolution within nitrogen fixing systems. Symp Soc Gen Microbiol 24:263–242 Priester JH, Olson SG, Webb SM, Neu MP, Hersman LE, Holden PA (2006) Enhanced exopolymer production and chromium stabilization in Pseudomonas putida unsaturated biofilms. Appl Environ Microbiol 72:1988–1996 Pritt B, O’Brien L, Winn W (2007) Mucoid Pseudomonas in cystic fibrosis. Am J Clin Pathol 128(1):32–34 Qiu D, Eisinger VM, Head NE, Pier GB, Yu HD (2008) ClpXP proteases positively regulate alginate overexpression and mucoid conversion in Pseudomonas aeruginosa. Microbiology 154(7):2119–2130 Remminghorst U, Rehm BH (2006) In vitro alginate polymerization and the functional role of Alg8 in alginate production by Pseudomonas aeruginosa. Appl Environ Microbiol 72(1):298–305 Reyes C, Peña C, Galindo E (2003) Reproducing shake flasks performance in stirred fermentors: production of alginates by Azotobacter vinelandii. J Biotechnol 105:189–198 Sabra WA (1998) Microaerophilic production of alginate by Azotobacter vinelandii. Dissertation, Technische Universität Carolo-Wilhelmina zu Braunschweig, Braunschweig, http://www.biblio.tu-bs.de/ediss/data/19990415a/19990415a.html Sabra WA, Hassan M (2008) Exopolysaccharide yield as a kinetic parameter for the statistical optimization of EPS production by Klebsiella pneumoniae biotechnology. ASCI 7(1):27–34 Sabra WA, Zeng A-P, Sabry S, Omar S, Deckwer W-D (1999) Effect of phosphate and oxygen concentrations on alginate production and stoichiometry of metabolism of Azotobacter vinelandii under microaerobic conditions. Appl Microbiol Biotechnol 52:773–780 Sabra WA, Zeng A-P, Lünsdorf H, Deckwer W-D (2000) Function and variation of alginate production in Azotobacter vinelandii under nitrogen fixation conditions. Appl Environ Microbiol 66(9):4037–4044 Sabra WA, Zeng A-P, Deckwer W-D (2001) Bacterial alginate: Physiology, product quality and process aspects. Appl Microbiol Biotechnol 56:315–325 Sabra WA, Kim E-J, Zeng A-P (2002) Physiological responses of Pseudomonas aeruginosa PA01 to oxidative stress in controlled microaerobic and aerobic cultures. Microbiology 148:3195–3202 Sabra WA, Lünsdorf H, Zeng A-P (2003) Alterations in the formation of lipopolysaccharide and membrane vesicles on the surface of Pseudomonas aeruginosa PAO1 under oxygen stress conditions. Microbiology 149:2789–2795 Sato SS, Mukataka H, Kataoka, Takahashi J (1984) Effects of pressure and dissolved oxygen concentration on growth of Pseudomonas aeruginosa. J Ferment Technol 62(1):71–75 Saude N, Junter G-A (2002) Production and molecular weight characteristics of alginate from free and immobilized-cell cultures of Azotobacter vinelandii. Process Biochem 37:895–900 Saude N, Chèze-Lange H, Beunard D, Dhulster P, Guillochon D, Cazé A-M, Morcellet M, Junter GA (2002) Alginate production by Azotobacter vinelandii in a membrane bioreactor. Process Biochem 38:273–278 Savalgi V, Savalgi V (1992) Alginate production by Azotobacter vinelandii in batch culture. J Gen Appl Microbiol 38:641–645
Microbial Production of Alginates: Physiology and Process Aspects
173
Schenk A, Weingart H, Ullrich MS (2008) The alternative sigma factor AlgT, but not alginate synthesis, promotes in planta multiplication of Pseudomonas syringae pv. glycinea. Microbiology 154(2):413–421 Schurr MJ, Martin DW, Mudd MH, Deretic V (1994) Gene cluster controlling conversion to alginate-overproducing phenotype in Pseudomonas aeruginosa: functional analysis in a heterologous host and role in the instability of mucoidy. J Bacteriol 176(11):3375–3382 Seanez G, Pena C, Galindo E (2001) High CO2 affects alginate production and prevents polymer degradation in cultures of Azotobacter vinelandii. Enzyme Microbial Technol 29:535–540 Skjak-Braek G (1992) Alginate: biosynthesis and some structure function relationships relevant to biomedical and biotechnological applications. Biochem Soc Trans 20(1):27–33 Stehling EG, Silveira WD, Leite Dda S (2008) Study of biological characteristics of Pseudomonas aeruginosa strains isolated from patients with cystic fibrosis and from patients with extrapulmonary infections. Braz J Infect Dis 12(1):86–88 Steigedal M, Sletta H, Moreno S, Maerk M, Christensen BE, Bjerkan T, Ellingsen TE, Espìn G, Ertesvåg H, Valla S (2008) The Azotobacter vinelandii AlgE mannuronan C-5-epimerase family is essential for the in vivo control of alginate monomer composition and for functional cyst formation. Environ Microbiol 10(7):1760–1770 Suh IS, Schumpe A, Deckwer WD (1992) Xanthan production in bubble column and air- lift reactors. Biotechnol Bioeng 39(1):85–94 Trujillo-Roldan M, Pena C, Ramirez OT, Galindo E (2001) Effect of oscillating dissolved oxygen tension on the production of alginate by Azotobacter vinelandii. Biotechnol Prog 17:1042–1048 Trujillo-Roldán MA, Peña C, Galindo E (2003) Components in the inoculum determine the kinetics of Azotobacter vinelandii cultures and the molecular weight of its alginate. Biotechnol Lett 25(15):1251 Wagner VE, Iglewski BH (2008) P. aeruginosa Biofilms in CF infection. Clin Rev Allergy Immunol 35:124–134 Wood LF, Leech AJ, Ohman DE (2006) Cell wall-inhibitory antibiotics activate the alginate biosynthesis operon in Pseudomonas aeruginosa: roles of sigma (AlgT) and the AlgW and Prc proteases. Mol Microbiol 62(2):412–426 Xu KD, Stewart PS, Xia F, Huang C-T, McFeters GA (1998) Spatial physiological heteroginicity in Pseudomonas aeruginosa biofilm is determined by oxygen availability. Appl Environ Microbiol 64(10):4035–4039 Young JM, Park DC (2007) Probable synonymy of the nitrogen-fixing genus Azotobacter and the genus Pseudomonas. Int J Syst Evol Microbiol 57(12):2894–2901 Young M (1983) Microbial polysaccharide. In comprehensive biotechnology, vol 3. Pergamon, Toronto, pp 1005–1044 Yuan LL, Li YQ, Wang Y, Zhang XH, Xu YQ (2008) Optimization of critical medium components using response surface methodology for phenazine-1-carboxylic acid production by Pseudomonas sp. M-18Q. J Biosci Bioeng 105(3):232–237 Zhang ZY, Zhong JJ (2004) Scale-up of centrifugal impeller bioreactor for hyperproduction of ginseng saponin and polysaccharide by high-density cultivation of panax notoginseng cells. Biotechnol Prog 20(4):1076–1081
Alginate-Based Blends and Nano/Microbeads Kamalesh Prasad and Jun-ichi Kadokawa
Contents 1 2
Introduction........................................................................................................................ Alginate Blends.................................................................................................................. 2.1 General Methods for the Preparation of Alginate Blends......................................... 2.2 Applications of Alginate Blends ............................................................................... 3 Alginate Microbeads and Nanobeads ................................................................................ 3.1 Methods for Preparation of the Beads ...................................................................... 3.2 Application of Alginate Beads .................................................................................. 4 Conclusion ......................................................................................................................... References ................................................................................................................................
176 177 177 178 183 183 188 203 203
Abstract Usually alginates are the sodium or calcium salts of alginic acid. Alginic acid is extracted from certain species of brown seaweeds. Sodium alginate is watersoluble, while calcium alginate is water-insoluble. Apart from their sourcing from the renewable resources in the nature, they have the added advantages of biocompatibility, biodegradability and ability to form a gel, which makes them suitable candidates for various applications. Alginates are mainly processed as capsules, beads, fibres and blends with other natural and synthetic polymers and are widely used in various fields. Developments made in the field of alginate-based blends and beads in recent years are summarized here. The method of preparation of the micro- and nano-sized beads, blends and their important applications are discussed, focusing on their recent uses in membrane science, drug slow release, biomedicine, process technology including enzyme immobilization, and pollution control. Recent applications of the beads in stem cell research are also briefly discussed.
K. Prasad Discipline of Marine Biotechnology & Ecology, Central Salt and Marine Chemicals Research Institute, Gijubhai Badheka Marg, Bhavnagar, 364 002, Gujarat, India J. Kadokawa() Department of Nano-structured & Advanced Materials, Graduate School of Science and Engineering, Kagoshima University, 1–21–40 Korimoto, Kagoshima, 890-0065, Japan e-mail: [email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_8, © Springer-Verlag Berlin Heidelberg 2009
175
176
1
K. Prasad and J. Kadokawa
Introduction
Alginates are typically the sodium and calcium salt derivatives of alginic acid, extracted from some species of brown algae. Although initially alginic acid was thought to consist of a uniform polymer of mannuronic acid, later studies showed the presence of both guluronic and mannuronic acid blocks. More precisely, it consists of (1→4)-linked b-d-mannuronic acid and a-l-guluronic acid (Fischer and Dorfel 1955). In alginic acid, both residues exist in the form of a homopolymer block or a heteropolymer block as shown in Fig. 1 (Haug et al. 1967). Because alginates are extracted from certain brown seaweeds, depending on the seaweed source and growing conditions the ratio of mannuronic acid to guluronic acid can vary in various samples of alginates. They have reversible gel-forming ability in the
HOOC O
HO O
HO
O
O
HO
O OH
HOOC M
OH O
O OH COOH
M
M OH
HOOC
OH O O
OH
OH
O
O
O
O OH
O HOOC
OH
G
G
HOOC G OH
HOOC
M
OH O
O OH O
O O
O HO
HOOC
HOOC G
O
HO OH
M
M= b-D - mannuronic acid ; G = a- L- guluronic acid
Fig. 1 Repeating d-mannuronic, l-guluronic and b-d-mannuronic-a-l-guluronic acid blocks in alginic acid
Alginate-Based Blends and Nano/Microbeads
177
presence of calcium. It has been found that the poly(guluronic acid) blocks can bind calcium ions more effectively than the poly(mannuronic acid) blocks (Grant et al. 1973). Owing to their biocompatible nature, alginates are well adapted for tissue engineering, as delivery vehicles for drugs and cells, for enzyme entrapment as well as in biomedicine as wound-dressing material (Sussman 2006). Apart from these applications, they are also used as chelators for pulling radioactive toxins such as iodine-131 and strontium-90 from the body and as food additives to provide structures to some food materials, including marine and freshwater products (http;// www.wikipedia.org; Llaneras 2000). Alginates can be processed as capsules, beads and fibres, blends with other natural and synthetic polymers or films. Some alginate blends and films are commercialized as haemostatic wound-dressing materials. Researchers are continuing their search for suitable materials useful in many newer applications, and consequently beads and blends obtainable from alginic acid and alginates are being successfully used in a variety of applications. On the basis of the above viewpoints, the recent advances made in the research related to beads and blends are reviewed in this chapter. A simple representative structure of the alginates will be used throughout the chapter.
2
Alginate Blends
Like many other natural polymers, alginates are also used for preparation of blends by conjugating them with other compatible synthetic and natural polymers or materials. The blends thus obtained show the properties of both the ingredients and sometimes such blends have properties superior to those of many other biopolymer blends and are useful for many applications.
2.1
General Methods for the Preparation of Alginate Blends
Cross-linking of alginates with suitable materials is the most popular method for the preparation of alginate blends. There are numerous examples of blends prepared by this technique. A blend composed of sodium alginate and hydroxyethylcellulose (HEC) ionically cross-linked with phosphoric acid was prepared by Kalyani et al. (2006) and its potential use as an evaporation membrane was tested and will be discussed later in this chapter. Blends of sodium alginate with other natural polymers are also found to have improved properties compared with the properties of the individual constituents. Among the natural polymers, chitosan has been widely used for preparation of such blends as the cross-linking is easier owing to presence of amino groups. Sodium alginate–gelatin and chitosan–sodium alginate biopolymeric blend materials cross-linked with glutaraldehyde were prepared by Fadnavis et al. (2003). Xu et al. (2007) have demonstrated the preparation of alginate-chitosan blend gel beads based on Ca2+ or dual cross-linking with various
178
K. Prasad and J. Kadokawa
proportions of alginate and chitosan. Another seaweed-based polysaccharide, carrageenan, was also blended with sodium alginate via cross-linking with the water-soluble carbodiimide moiety to prepare a membrane (Xu et al. 2003) and details will be discussed later. Simple blending or mixing was also found to be successful for the preparation of alginate blends. Blending of 84% deacetylated chitosan and sodium alginate biopolymers by cost-effective and simple fabrication techniques for direct methanol fuel cell application was investigated by Smitha et al. (2005). To introduce the antimicrobial property of chitosan into alginate, Ariyakriangkrai et al. (2005) blended sodium alginate with O-carboxymethyl chitosan or N-(carboxyacyl) chitosan. More recently, a spinning technique was employed for the preparation of effective blends of alginates. Wang et al. (2006) reported the preparation of alginate and soy protein isolate blend fibres by spinning their solution through a viscose-type spinneret into a novel coagulating bath containing aqueous CaCl2, HCl and ethanol. Alginate and gelatin blend fibres were also prepared by a spinning technique in the same manner (Fan et al. 2005). The same researchers have also reported the preparation of alginate-carboxymethyl chitosan blend fibres by this technique (Fan et al. 2006). They showed a good miscibility between alginate and carboxymethyl chitosan owing to the existence of a strong interaction from the intermolecular hydrogen bonds. The introduction of carboxymethyl chitosan into the blend fibre was found to improve the water-retention properties of the material in comparison with those of pure alginate fibre.
2.2
Applications of Alginate Blends
2.2.1
Membrane Science
Owing to the excellent film-forming ability of the alginates as well as that of their blends, their performances in thin-film form are being widely investigated by many researchers as potential membranes. As a result of the hydrophilic nature of sodium alginate, its applications in the separation of non-aqueous solvent mixtures by the perevaporation technique have been more successful. Hence attempts have been made by many researchers to introduce hydrophobicity into the sodium alginates by reacting or blending them with suitable hydrophobic substrates. Improved membrane performance of sodium alginate blended with HEC over sodium alginate film for 2-propanol dehydration was reported by Naidu and Aminabhavi (2005). They further found the highest selectivity for water using such membranes by incorporating zeolite [ZSM-5(40)] in the membrane, which showed enhanced flux. Further, the same group has reported the compatibility of sodium alginate and HEC and perevaporation separation of water + 1,4-dioxane and water + tetrahydrofuran (Naidu et al. 2005a, Naidu et al 2005b). The performance of the blend membranes of sodium alginate and HEC, ionically cross-linked with phosphoric acid for perevaporation separation of a t-butanol-water mixture, was evaluated by Kalyani et al. (2006). The blended polymer thus obtained was found to have good potential for breaking
Alginate-Based Blends and Nano/Microbeads
179
the azeotrope of 88 wt% t-butanol with high selectivity. Yang et al. (2000) studied the perevaporation separation of ethanol–water mixtures through blend membranes prepared with 8 wt% cellulose cuoxam and 8 wt% aqueous sodium alginate solution cross-linked by using a Ca2C bridge in 5 wt% CaCl2 aqueous solution. They studied the effect of Ca2C on the flux and other parameters. A hollow-fibre composite membrane made up of a poly(vinyl alcohol)–sodium alginate blend along with maleic acid, supported by a polysulfone hollow-fibre ultrafiltration membrane was prepared to study the perevaporation dehydration from 2-proanol, butanol, t-butanol and ethanol aqueous solutions. The results of the perevaporation experiment showed high selectivity and promising permeability for the blended membrane cross-linked with 1.5 wt% maleic acid (Dong et al. 2006a). Apart from perevaporation applications of the alginate blend membranes, sodium alginate–poly(vinyl pyrrolidone) blend membranes were used for the separation of aqueous dimethylformamide solutions in the concentration range 0–100 wt% (Solak et al. 2008). Phisalaphong et al. (2008) have used bacterial cellulose and alginate in an aqueous NaOH–urea solution as substrate materials for the fabrication of a novel blend membrane using supercritical CO2 drying. The supercritical CO2 drying provided a nanoporous structure that had better water adsorption capacity and water vapour transmission rate than the conventional blend membranes prepared without using supercritical CO2 drying. Poly(ion complex) membranes made by blending 84% deacetylated chitosan and sodium alginate biopolymers followed by cross-linking with glutaraldehyde were tested for the separation of ethanol-water mixtures. These blend membranes (Fig. 2) were found to have good potential for breaking the azeotrope of water (Kanti et al. 2004).
2.2.2
Slow-Release Applications
Owing to the biocompatible nature of the alginates and their blends with natural polymers, they are being investigated as potential candidates for various drug slowrelease applications. Chitosan–alginate blend microcapsules were evaluated as a method for oral delivery of egg yolk immunoglobulin Y antibodies. It was found that the stability of immunoglobulin Y in simulated gastric fluid was greatly improved by encapsulation in chitosan–alginate blend microcapsules (B. Li et al. 2007). Sarmento et al. (2007) have demonstrated that nanobeads obtained from an alginate–chitosan blend were effective for oral insulin delivery for rats. Alginatechitosan blend gel beads based on Ca2+ prepared by dual cross-linking with various proportions of alginate and chitosan as stated earlier were used for the sustained release of bovine serum albumin, a model protein drug in simulated gastric, intestinal and colonic fluid by Xu et al. (2007). It was observed that the dual cross-linking promoted the stability of the beads under gastrointestinal tract conditions. Films of alginate and gelatin, cross-linked with Ca2+, with ciprofloxacin hydrochloride as a model drug incorporated in different concentrations were obtained by Dong et al. (2006b). Slow-release profiling of an anti-inflammatory drug (diclofenac sodium) from the polymeric blend beads of poly(vinyl alcohol)-g-polyacrylamide with
180
K. Prasad and J. Kadokawa O
ONa
C O
ONa
O
OH
C
OH O
HO
O
OH
O
O
O
O
HO
C
O
OH
ONa
O OH
C
O
O
ONa NH2 HO
O O HO
O
O O CH2OH
NH2 C
H
NaO
O
Glutaraldehyde
CH2OH
O
H
C
O
NaO
C O
O
C
OH O
O
O
OH
O
O
HO
O
O
C
O
OH
O
O O ONa
O
C
OH O
H C
O
ONa CH2OH
H
CH2
+ H3N HO
O O
O
O
O
O H2C
N CH
O O
C
H
C
H
Fig. 2 Formation of blend films of sodium alginate and chitosan cross-linked with glutaraldehyde
sodium alginate cross-linked with glutaraldehyde was studied by Sanli et al. (2007), and they observed a decrease in the rate of drug release with increasing poly(vinyl alcohol) to sodium alginate ratio, drug to polymer ratio and extent of cross-linking. Polymeric blend microspheres consisting of sodium alginate and methylcellulose were prepared by a water-in-oil emulsion method followed by cross-linking with glutaraldehyde and loaded with nifedipine, an anti-inflammatory drug, and the release characteristics of the drug were evaluated by Ramesh Babu et al. (2007). Poly (dl-lactic acid) was combined with calcium alginate fibres using supercritical
Alginate-Based Blends and Nano/Microbeads
181
CO2 to form dual protein release scaffolds and subsequently more protein release from the hydrophilic alginate fibres was observed by Ginty et al. (2008). A blend of sodium alginate and hydroxypropyl methylcellulose was used as the candidate for modulating barrier material for pulse drug release by B. Li et al. (2008). The results revealed that programmed drug delivery to achieve pulsatile drug release three times daily can be obtained from these blend tablets in a capsule system by a systemic formulation approach. Almeida and Almeida (2004) reported preparation of pindolol (a pulsatile drug)-loaded alginate–gelatin beads using a solvent-free technique that involves a cross-linking reaction. Composite beads consisting of xanthan, a branched biopolymer, and sodium alginate were prepared using an ionotropic gelation method by Pongjanyakul and Puttipipatkhachorn (2007) and subsequently studies on the entrapment efficiency of diclofenac sodium, the thermal properties, water uptake, swelling and release in various media were carried out. They also established the molecular interactions between xanthan and sodium alginate, which modulates the physicochemical properties and drug release. Guar gum was also used to prepare blends with alginates. A pH-sensitive alginate–guar gum hydrogel cross-linked with glutaraldehyde was designed for the controlled delivery of protein drugs. The presence of guar gum and glutaraldehyde cross-linking increases the entrapment efficiency and prevents the rapid dissolution of alginate in the high pH of the intestine, ensuring a controlled release of the entrapped drug (George and Abraham 2007). Sodium alginate–carboxymethyl guar gum biopolymeric beads cross-linked with Ba2+ were prepared with the intention of achieving greater stability in the gastrointestinal tract (Bajpai et al. 2006).
2.2.3
Miscellaneous Applications of Alginate Blends, Including Biomedical Applications
The ideal cell-carrier material for cartilage regeneration should be one that closely mimics the natural environment in a living articular cartilage matrix. Alginatebased chitosan hybrid biomaterials could provide excellent supports for chondrocyte adhesion. To test this assumption, Iwasaki et al. (2004) investigated the behaviour of adhesion of rabbit chondrocytes onto an alginate polymer compared with the behaviour of adhesion of the chondrocytes onto some alginate-based chitosan hybrid polymer fibres in vitro. It was found that the alginate-based chitosan hybrid polymer fibres had much improved adhesion capacity with chondrocytes in comparison with alginate polymer fibre alone, showing its viability as a desirable material for cartilage tissue scaffolds. Rinaudo (2008) recently summarized some important biomedical applications of alginate blends. One of the applications described in the review involved the thermosensitive gel of calcium alginate and poly(N-isopropylacrylamide) (PNIPAAm). The characteristic lower critical solution temperature of this blend was tailored to be close to human body temperature by controlling the amount of N-isopropylacrylamide and calcium alginate during the synthesis. New biomaterial for surgical applications based on viscoelastic blends of alginate–hyaluronate that merge the gel-forming properties of alginate
182
K. Prasad and J. Kadokawa
with the healing properties of hyaluronate was also proposed. Another biomaterial discussed was the self-curing hydroxyapatite/alginate mixture, which has the potential for use in plastic and reconstructive as well as in oral and maxillofacial surgery. Targeting to develop biologically active scaffolds from polysaccharide mixtures, Donati et al. (2007) reported the preparation of alginate and a lactosemodified chitosan mixture under physiological conditions and at a semidilute concentration avoiding associative phase separation. A series of blended films based on poly(ethylene oxide) and sodium alginate were prepared by a solution casting method (Fig. 3). These blends were found to have mechanical properties superior to those of poly(ethylene oxide) films alone (Caykara et al. 2005). Chitosan– alginate–hyaluronate complexes modified by an arginine–glycine–aspartic acidcontaining protein as tissue-engineering scaffolds for cartilage regeneration was evaluated by Hsu et al. (2004). Alginate blends were also found to be suitable for artificial organ manufacturing, e.g. bioartificial pancreas (BAP), which is a medical device enclosing insulinsecreting cells in a semipermeable membrane that is expected to physiologically control the glucose levels and thus to be able to inhibit development of serious chronic complications in diabetic patients. Sakai et al. (2002) utilized a sol–gel reaction to prepare an immunoisolatable membrane for a microcapsule-shaped bioartificial pancreas derived from two precursors, 3-(aminopropyl)trimethoxysilane and tetramethoxysilane, formed on calcium alginate gel beads via electrostatic
CH2CH2
O
CH2CH2
O
n
n
ONa
O C O
O
G
H G O O
O O OH
O
OH
C
O HO
ONa H O O M
HO O O
H O
O C
O O
H M ONa
C
CH2CH2
ONa CH2CH2
O
O
n n
Fig. 3 Blends based on poly(ethylene oxide) and sodium alginate
Alginate-Based Blends and Nano/Microbeads
183
Islets Islets Oxygen, nutrient
Insulin, wastes
Islets Islets
Membrane made up of alginate blends
Fig. 4 Layout of bioartificial pancreas
interaction. Iwata et al. (2004) also demonstrated the preparation of a bioartificial pancreas from alginate blends (Fig. 4). a-Galactosidase from Aspergillus oryzae was effectively immobilized in gelatinblended alginate hydrogel fibres and glycerol hardened by cross-linking with glutaraldehyde. It was observed that thus immobilized a-galactosidase was more stable at higher temperature and lower pH (Naganagouda and Mulimani 2006). A strong hydrogel system based on the blends of polyacrylamide, agar and sodium alginate with water absorbency and pH resistance superior to those of their unblended counterparts was developed by Meena et al. (2008). Roy et al. (2005) utilized crosslinked alginate–guar gum beads to fluidize the bed affinity media for purification of jacalin (a lectin present in the seeds of the jackfruit). The preparation of an alginate– polymethacrylate hybrid material by radical polymerization of the cationic methacrylate monomer (2-aminoethyl methacrylate hydrochloride) in the presence of sodium alginate in water was described by Kadokawa et al. (2005). The insoluble material was formed with progress of the polymerization as shown in Fig. 5.
3 3.1
Alginate Microbeads and Nanobeads Methods for Preparation of the Beads
Alginate beads have been popularly used for the last several decades for immobilization of enzymes and were prepared conventionally by gelation of sodium alginate solution in the presence of calcium salt (Fig. 6). The beads obtainable by this technique are normally large and their sizes are controlled by the orifice diameter
+
OH
OH O
+
+
AEMA
lonic complex
COO Na Sodium alginate
–
O HO
–
COO Na OH O HO O
AEMA = 2-aminoethyl methacrylate
O
O OH
OH
+
Alginate K2S2O8
+
OH
OH O
lonic bond
Radical polymerization
–
O OH COO Na
O
–
COO Na
O
+ +
–
PolyAEMA
AEMA-HCl
O
COCH2CH2NH3 Cl
CH3
Alginate-polyAEMA blend
H
H
Fig. 5 Preparation of hybrid material by polymerization of 2-aminoethyl methacrylate hydrochloride in the presence of sodium alginate
O
–
COO Na
184 K. Prasad and J. Kadokawa
Alginate-Based Blends and Nano/Microbeads
185
–O C Ca 2+
O
Ca2+
COONa+ O
C
O–
Ca-alginate bead
Fig. 6 Basic mechanism for the formation of calcium alginate beads
employed in their preparation. The diameters of the beads thus obtained were generally 2–5 mm (Mofidi et al. 2000). A schematic diagram of the conventional bead-forming apparatus is depicted in Fig. 7. Owing to the demand for micro- and nano-sized beads for a variety of applications, the methods for the preparation of the beads were improved by employing various techniques. For generation of beads with smaller diameter, Zhao et al. (2007) developed an electrodispersion reactor to prepare calcium alginate microgel beads suitable for microencapsulation. In the electrodispersion reactor, pulsed electric fields were utilized to atomize aqueous mixtures of sodium alginate and CaCO3 nanoparticles (dispersed phase) from a nozzle into an immiscible insulating second liquid (continuous phase) containing a soluble organic acid. The sizes of the beads obtained by this technique changed from 412 ± 90 to 10 ± 3 mm as the applied peak voltage was increased (Fig. 8). Calcium alginate–silica dehydrogenase hybrid biocompatible beads were prepared by the polymerization of silica particles. The catalytic activity of the enzyme entrapped in such beads was investigated by Xu et al. (2006). The catalytic activity of yeast alcohol dehydrogenase entrapped in alginate-silica beads was found to be superior to that of yeast alcohol dehydrogenase entrapped only in alginate beads. The schematic layout for such bead formation is depicted in Fig. 9. Production of small and monodispersed alginate beads (diameter 0.5–1 mm) for cell immobilization by destabilizing a viscous jet with a controlled disturbance technique in a bioreactor for high-density operations without the difficulties of cell separation was described by Seifert and Phillips (1997). Nedovic et al. (2001) were also successful in preparing alginate microbeads of diameter 250 mm–2.0 mm loaded with brewing yeast by an electrostatic technique. Laminar jet break-up technology was also utilized to prepare sodium alginate microbeads by Del et al. (2005). Alginate beads with ionic properties were synthesized through the conjugation reaction of alginate and tyramine (a monoamine compound derived from the amino acid tyrosine) by immersing an aqueous solution of the alginate containing horseradish peroxidase into a solution containing H2O2, caused the solution to gel
186
K. Prasad and J. Kadokawa Syringe containing sodium alginate solution
CaCl2 solution
Magnetic stirrer
Fig. 7 Conventional apparatus used for the preparation of alginate beads
Na-alginate + CaCO3 nano particles (dispersed phase)
Nozzle Insulating liquid
High voltage power
Continuous phase
collector
Ca-alginate micro beads
Fig. 8 Generation of calcium alginate beads by an electrodispersion technique
via peroxidase-catalysed oxidative coupling of the phenols as shown in Fig. 10 (Sakai and Kawakami 2007). Owing to the importance of small beads for specific uses in biomedicine and bioengineering, Sugiura et al. (2005) utilized a novel microfluidic device, where a silicon micronozzle array was used to produce 50–200-mm calcium alginate beads
Alginate-Based Blends and Nano/Microbeads COONa+
187 COONa+
COONa+
COONa+
Ca2+
Polymerization
COONa+ COONa+
COONa+
COONa+
COONa+
COONa+
Ca-alginate biocomposite beads with polymerized silica.
Na-alginate containing enzyme
= Enzyme
COONa+ COONa+
= Alginate
= Silica
Fig. 9 Preparation of alginate beads by an in situ polymerization technique
COO–
HN
H2N OH
+
EDC /NHS
C
O
OH
pH =6.0
Ca-alginate bead
Tyramine
HRP /H2O2
Alginate bead with tyramine moiety Conjugated tyramine moiety
Ca-alginate ionic conjugated gel bead EDC = 1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide ; NHS = hydroxysulfosuccinimide ; HRP = Horseradish peroxidase
Fig. 10 Synthesis of tyramine-conjugated alginate gel beads
with narrow size distribution. They were successful in encapsulating living cells in such small calcium alginate beads. Continuous generation of hydrogel beads smaller than 100 mm using a double T-junction in a microfluidic channel for alternate injection of aqueous fluids inside a droplet unit carried within immiscible oil was obtained by Um et al. (2008). They encapsulated biological materials in such beads using a microfluidic droplet-merging channel. A novel polymer gel having characteristic temperature and magnetic field sensitivity was prepared by incorporating magnetite nanoparticles into PNIPAAm hydrogels synthesized in calcium alginate beads (Xulu et al. 2000). The enzymatically modified alginate beads prepared by Morch et al. (2007) were found to be more elastic, compact, less permeable and extremely stable under physiological conditions, offering significant advantages over native alginates. A chemoenzymatic reaction was used to prepare tailor-make alginate-bearing galactose beads by the reaction of 1-amino-1-deoxygalactose with
188
K. Prasad and J. Kadokawa
a mannuronan sample via an N-glycosidic bond involving the carboxyl group of the mannuronic acid residues, resulting in the formation of modified epimerized material with the ability to form stable gels (Donati et al. 2005). Alginate–glycerol–chitin beads with improved antifungal activity were synthesized by Zohar–Perez et al. (2004). Kim et al. (2007) demonstrated the preparation of alginate-quaternary ammonium complex beads with antimicrobial activity by the reaction of sodium alginate with 3-(trimethoxysilyl)propyl octadecyldimethylammonium chloride (TSA) in acid solution. The sodium alginate–TSA complex beads obtained were a new type of insoluble cationic polymer which can kill or remove microorganisms in water by mere contact without releasing the reactive agent. Covalently linked alginate beads with high stability and cell compatibility were achieved by grafting of mannuronan (alginate with 100% mannuronate) with methacrylate moieties by photoinitiated polymerization in the alginate beads (Rokstad et al. 2006). The stability and the cell compatibility of the beads thus obtained may make them attractive as bioreactors for delivering therapeutic proteins.
3.2
Application of Alginate Beads
3.2.1
Process Technology Employing Immobilization on Alginate Beads
Processes or reactions yielding enantiomeric pure compounds are very important in pharmaceutical industries. Many reactions catalysed by enzymes immobilized in alginates beads have been found to provide such pure compounds. The asymmetric bioreduction of ethyl 3-halo-2-oxo-4-phenyl butanoate catalysed by several microorganisms, especially by Saccharomyces cerevisiae immobilized in calcium alginate beads with a double gel layer, was carried out by Milagre et al. (2006) (Fig. 11). They reported the highest chemical yield for the above-mentioned bioreduction as 90%, with 70% diastereomeric excess and 96–99% enantiomeric excess. They further stated that the entrapment of cells with double gel layers was fundamental to achieve high enantioselectivities and diastereoselectivities. Hann et al. (2002) reported Acidovorax facilis 72W nitrilase catalysed hydrolysis of 2-methylglutaronitrile to 4-cyanopentanoic acid (as the ammonium salt), an intermediate in the preparation of 1,5-dimethyl-2-piperidone, an important commercial
O
Saccharomyces cerevisiae immobilized in Ca-Alginate Beads
O O
O X
O
X
Ethyl 3-halo-2-oxo-4-phenyl butanoate
Fig. 11 Bioreduction of ethyl 3-halo-2-oxo-4-phenylbutanoate
OH
O
+
O X
OH
Alginate-Based Blends and Nano/Microbeads
189
precision cleaning solvent (Fig. 12). The immobilization of the enzyme cells was done in alginate beads instead of carrageenan gel followed by cross-linking with glutaraldehyde and polyethyleneimine. The catalyst thus formed was stable in reaction mixtures containing high concentrations of 4-cyanopentanoic acid ammonium salt. Evidence for Fenton photoassisted processes mediated by encapsulated Fe ions in alginate beads at biocompatible pH was demonstrated by Fernandez et al. (2000). They stated that the iron carboxylate complex was active during the Fentonenhanced decolouration/degradation of orange II via the encapsulated Fe catalyst. Baker’s yeast cells entrapped in alginate beads were shown to catalyse reactions in organic solvents when a cofactor regeneration scheme was implemented. Reduction of acetophenone to 1-phenylethanol, a chiral alcohol with high enantiomeric excess using baker’s yeast as well as a cosubstrate to regenerate the cofactor was reported by Griffin et al.(1998) (Fig. 13). Owing to its ability to preserve natural colours in juice and health drinks, l-malic acid is considered to be very important in the food processing industry and thus various methods to prepare pure l-malic acid have been attempted by many researchers. Bressler et al. (2002) converted fumaric acid to l-malic acid in a bioreactor catalysed by S. cerevisiae immobilized in small glasslike beads of an alginate-silicate sol-gel matrix (Fig. 14). The conversion reported by them was almost 100%, which was above the equilibrium value of approximately 84% and higher than that of the industrial process. They were also able to obtain pure l-malic acid in contrast to the existing industrial biocatalytic process, which yields l-malic acid salts. Gervais et al. (2000) studied the effect of aeration during cell growth on the subsequent reduction of 2-hexanone and 2-octanone by yeast cells entrapped in calcium alginate beads. They used 2-propanol as a sacrificial substrate to regenerate the cofactor NAD(H). The use of strictly aerobic conditions when growing the cells resulted in the highest initial reaction rates, as well as the production of only a single product, i.e. the enantiomeric excess of the (S)-alcohols was 100% (Fig. 15). Some more important chemical reactions catalysed by enzymes entrapped in alginate beads have been studied. Hydrolysis of starch was carried out by a blend of individually entrapped glucoamylase and pullulanase in calcium alginate beads in a ratio of 3:2 (activitywise), the individually entrapped enzymes showed enhanced thermostability at 55°C and effective hydrolysis was observed (Roy and Gupta 2004). Won et al. (2005) entrapped lipase from Candida rugosa by dropwise addition of an aqueous mixture of sodium alginate and putting the biocatalyst into a hardening solution of a Ca2+ salt. They observed a higher rate of reusability of the catalyst coated with alginate beads in comparison with the non-coated ones. Yadav and Jadhav (2005) preimmobilized the biocatalyst Candida antarctica lipase B in hexagonal mesoporous silica followed by encapsulation in calcium alginate beads for the transesterification of p-chlorobenzyl alcohol with vinyl acetate to give p-chlorobenzyl acetate at 30°C. They found excellent reusability of the catalyst with a decrease of only 4% in the overall conversion of the transesterification reaction even after the fourth reuse. A microbial bioreactor based on calcium alginate immobilized Lactobacillus cells coupled to a pH electrode was developed for quantitative estimation of carbohydrate fermentation activity. The procedure developed
2-Methylglutaronitrile
Fig. 12 Nitrilase-catalysed hydrolysis of 2-methylglutaronitrile
N N
N A. facilis 72W nitrilase immobilised in alginate bead
4-Cyano pentanoic acid (Ammonium salt)
O
–
+
ONH4
Pd/ C, H2
CH3NH2 N
1,5 Dimethyl piperidone
O
190 K. Prasad and J. Kadokawa
Alginate-Based Blends and Nano/Microbeads O
191 OH
Substrate reduction
OH
Baker’s yeast entrapped in alginate bead
+ + NADH Alcohol NAD deydragenase Acetophenone
+H+
1-Phenylethanol (S)
1-Phenylethanol (R)
OH
O
2-Hexanone
Co-Substrate oxidation
2-Hexanol
Fig. 13 Reduction of acetophenone by baker’s yeast immobilized in alginate beads in the presence of a cosubstrate, 2-hexanol
O O OH
OH Saccharomyces cerevisiae HO immobilised on alginate bead
HO
OH
O
O
L-Malic acid
Fumaric acid 100 % conversion to free L-malic acid
Fig. 14 Conversion of fumaric acid to l-malic acid
was simple, being based on pH variation that can give quantitative results, in comparison with other existing techniques for the carbohydrate fermentation pattern from which only qualitative results could be obtained. It also offers reduction in time and costs and is a suitable tool for the rapid analysis of isolated strains and in studies of modifications of sugar metabolism in mutants (Corton et al. 2000). An immobilization culture of Nicotiana tabacum cells on calcium alginate beads to produce scopoletin (a naturally occurring coumarin) was achieved by adding coconut oil as an extractive solvent (Iizuka et al. 2005). d-Tagatose (an epimer of d-fructose) was continuously produced using thermostable l-arabinose isomerase immobilized in alginate beads with d-galactose solution in a packed-bed bioreactor (Ryu et al. 2003). Production of a versatile chiral building block, (R)-(−)-mandelic acid, by dynamic resolution of mandilonitrile by nitrilase from Alcaligenes faecalis immobilized in alginate beads was described by Kaul et al. (2006) (Fig. 16). They reported reusability of the biocatalyst up to 35 batches by the immobilization as compared with nine batches for free cells, and cross-linking extended it further to 40 batches. Acetophenone, a useful solvent for many chemical syntheses and an
192
K. Prasad and J. Kadokawa OH O
Substrate reduction
OH
+ 2-Hexanol (R)
2-Hexanol (S)
2-Hexanone or
or NADH Alcohol NAD+ +H+ deydragenase O 2-Octanone
+ OH
OH 2-Octanol (R)
2-Octanol (S)
HO
O Co-substrate oxidation
2-Propanol
2-Propanone
Fig. 15 Reduction of 2-hexanone or 2-octanone to the corresponding alcohols in the presence of a cosubstrate, 2-propanol N
N R-selective nitrilase Alcaligenes faecalis
OH
HO
Immobilised on alginate bead
(RS)-Mandelonitrile
COOH
+ H
(S)-Mandelonitrile
HO
H (R)-(–)-Mandelic acid
O + HCN
Benzaldehyde
Fig. 16 Dynamic resolution of mandilonitrile by nitrilase
important chemical in the perfume industry, was obtained by the transamination of (S)-(−)-a-methylbenzylamine by (S)-aminotransferase immobilized in alginate beads by Martin et al. (2007) (Fig. 17). Ascorbic acid is a very important antioxidant and plays a very crucial role in human metabolism. However, it has a tendency for oxidation, causing its degradation, and hence to stop the oxidation, suitable chemical modification of the structure of the antioxidant was performed by the effective utilization of some biocatalysts. Hsieh et al. (2007) used mycelium-bound enzymes of Aspergillus niger as a biocatalyst
Alginate-Based Blends and Nano/Microbeads
193 O
O NH2 O CH3 + H3C
O
(S)-Aminotransferase
OH CH3 +
Immobilised on Ca-alginate beads
OH
NH2
Pyruvic acid (S)-(–)-a-Methylbenzylamine
Acetophenone
Alanine
Fig. 17 Production of acetophenone by transamination process
entrapped in calcium alginate beads for the glucosylation of ascorbic acid using maltose as an acyl donor to yield ascorbic acid glucoside. They found exceptional reusability and storage stability of the entrapped mycelia. They reported thr final yield of ascorbic acid glucoside as 75%. Xytitol, obtainable from xylose, is a commercially important sweetener and is widely used in food and pharmaceutical industries. Candida guilliermondii FTI 20037 cells were entrapped in calcium alginate beads used as a biocatalyst for xylose to xylitol bioconversions during five successive batches in a stirred tank reactor by Carvalho et al. (2008). A gramnegative, rod-shaped, aerobe, Alcaligenes faecalis, entrapped on calcium alginate beads was used to convert 2-propanol to acetone by Mohammad and Bustard (2008). Owing to the increased amount of research into the preparation of environmentallly friendly polymers, the biodegradable lactic acid based polymers have been drawing attention for a long time. Apart from such applications, lactic acid is also part of many commercial foodstuffs. Research is also going on into the production of lactic acid by various methods, including enzyme-catalysed ones. In this direction, Givry et al. (2008) reported the production of lactic acid by the consumption of hexoses/pentoses by Lactobacillus bifermentans immobilized in alginate beads in optimized culture medium. Hydrogen, a future fuel, was produced under anaerobic conditions from municipal sewage sludge immobilized on calcium alginate beads (Wu et al. 2002). Urease, the enzyme which hydrolyses urea to CO2 and ammonia, is very susceptible to chymotrypsin and other proteases. DeGroot and Neufeld (2001) encapsulated urease in alginate beads coated with chitosan, poly(l-lysine) or poly(methylene-co-guanidine) membrane and were able to substantially protect urease from these proteases.
3.2.2
Slow Release of Drugs and Agrochemicals
Alginate beads being biocompatible are widely used in drug slow-release applications. There have been many reports of such applications in the last decade. We discuss the developments in this field in recent years by mentioning the most important works. Calcium alginate, owing to the strong gel network, can hold drugs in the gel matrix and slowly releases them in suitable environments. Murata et al. (2000) prepared calcium alginate gel beads with vegetable oil or chitosan, capable of floating in the gastric cavity. They studied the behaviour of the release of metronidazole
194
K. Prasad and J. Kadokawa
(an anti-infective drug) entrapped in such beads into artificial gastric juice. They observed not only sustained release but also gastric-mucosa-specific drug release from both types of bead. Later, the same research group prepared alginate gel beads containing weak acid salts of chitosan and ascorbic acid having rapid bile acid uptake ability (Murata et al. (2002). An anti-hapten immunoglobulin G (antibody) was covalently immobilized on glutaraldehyde-activated alginate–chitosan gel beads by Albarghouthi et al. (2000). It was concluded that antibodies could be successfully immobilized on alginate–chitosan gel beads. Such systems can be applied for the development of immunoaffinity purification and immunoassays. New interpenetrating polymer networks of sodium alginate with gelatin or egg albumin cross-linked with glutaraldehyde beads were prepared by Kulkarni et al. (2001) and the in vitro release of the broad-spectrum antibiotic cefadroxil incorporated in such beads was studied. To study the release profile of sodium diclofenac, alginate–chitosan particles were prepared by ionic gelation (Ca2+ and Al3+) by Gonzalez-Rodriguez et al. (2002). The alginate to chitosan ratio and the nature of the gellifying cation were found to control the rate of release of the drug. The release profile of a model drug (brilliant blue) entrapped in chitosan-coated calcium–alginate gel beads was studied by Shu and Zhu (2002). Alginate–chitosan beads cross-linked with a natural crosslinker, genipin, were prepared by Mi et al. (2002) and the rate of release of indomethacin (an anti-inflammatory drug) from these beads was studied. It was observed that the rate of release of the drug from the beads increased with the decrease in pH or concentration of alginate in the gelling solution owing to the decrease in the crosslinking density of the beads. Furthermore, good cellular compatibility for the beads was established by cytotoxic examinations. Time-programmed release of the macromolecule dextran was achieved by Iskakov et al. (2002) by utilization of calcium alginate gel beads modified with coated copolymer layers by varying the coating thickness. Choi et al. (2002) demonstrated a novel technique for the preparation of floating alginate beads from a sodium alginate solution using CaCO3 or NaHCO3 as gas-forming agents and investigated the profile of the release of riboflavin from these floating beads. 5-Fluorouracil, an anticancer drug, encapsulated in alginate beads was prepared and the release characteristic of 5-fluorouracil were established by Arica et al. (2002). The influence of microwave irradiation on the drug release properties of alginate, alginate–chitosan and chitosan beads was studied and it was observed that the release-retarding properties of alginate and alginate–chitosan beads were significantly enhanced by subjecting the beads to microwave irradiation (Wong et al. 2002). Ueng et al. (2004) investigated the lyophilized poly(l-lysine)-coated alginate antibiotic delivery system in vivo for the treatment of musculoskeletal infections. Biodegradable beads with a uniform diameter of 3 mm were prepared by mixing sodium alginate with vancomycin, coated with poly(l-lysine), followed by lyophilization. This study offered a biodegradable delivery system for antibiotics to treat musculoskeletal infections. Calcium alginate encapsulated vascular endothelial growth factor, which is a potent signal transduction molecule that acts specifically on vascular endothelial cells, was prepared by Gu et al. (2004). They further investigated the release of the entrapped vascular endothelial growth factor in phosphate-buffered saline and serum media. Sodium polystyrene sulphonate is a medicine used to treat
Alginate-Based Blends and Nano/Microbeads
195
high levels of potassium in the body and is normally administered in conjunction with sorbitol. The release of this drug from calcium alginate hydrogel beads was studied by Rousseau et al. (2004) and it was found that the structure of the crosslinked calcium alginate network was important for the retention and/or release of the sodium polystyrene sulphonate. To establish the merit of multilayered alginate beads in certain slow-release applications, Anal and Stevens (2005) prepared multilayer alginate beads by mixing chitosan, Ca2+ (a cationic component) and sodium alginate, and polyphosphate (anionic component) in various proportions. The multilayered bead thus obtained showed 20–30% release of ampicillin in comparison with 70% release by single-layered alginate beads. Kim et al. (2005) prepared the beads of the graft copolymers of alginate and PNIPAAm by dropwise additions of the aqueous solutions of the copolymers in Ca2+ solutions. They further studied the release of blue dextran from such beads. It was observed that owing to the presence of a temperature-sensitive component in the beads, the beads showed a higher release rate at high temperature and a lower release rate at low temperature. Diclofenac calcium alginate beads containing glyceryl palmitostearate were prepared by an ionotropic gelation method. The effect of the amount of glyceryl palmitostearate and heat treatment on characteristics of the diclofenac calcium alginate beads was investigated by Pongjanyakul et al. (2006). Alginate-whey protein isolate microspheres of different whey protein isolate to alginate ratio, particle diameter and concentration of polymer-bead-forming solution were prepared to develop a biocompatible vehicle for oral administration of bioactive compounds by Chen and Subirade (2006). They carried out studies on the release of riboflavin from these beads in simulated gastric fluid and found a satisfactory retarding property for the drug. Swelling studies and in vitro release of verapamil (a drug used for hypertension) from calcium alginate and calcium alginate–chitosan beads were investigated by Pasparakis and Bouropoulos (2006) and 80% encapsulation of the drug in both calcium alginate and calcium alginate–chitosan beads was observed. They further observed retardation of the drug release in the presence of chitosan in wet beads. A general schematic approach of entrapment of an active agent (e.g. drug) in alginate beads and subsequent coating with another natural polymer is shown in Fig. 18. To design oral delivery vehicles for peptide or protein drugs, pH-sensitive hydrogels such as alginate and chitosan have attracted increasing attention. Most of the synthetic polymers are immunogenic, and the incorporation of proteins into these polymers requires a harsh environment which may denature and inactivate the desired protein. The favourable properties such as biocompatibility, biodegradability, pH sensitivity, mucoadhesiveness, etc. have enabled these natural polymers to become the choice of pharmacologists as oral delivery matrices for proteins (George and Abraham 2006). Emulsan is a polyanionic heteropolysaccharide bioemulsifier produced by Acinetobacter calcoaceticus RAG-1 and has the natural ability to bind protein. This nature of emulsan was utilized by Castro et al. (2005) to prepare emulsan–alginate beads to study triggered release of protein. They were successful in demonstrating the assay of protein release of azo-bovine serum albumin using lipase and subtilisin (a protease). A schematic diagram for protein release from emulsan–alginate beads using protease is shown in Fig. 19. Alginate beads were prepared in conjugation with
O
O
O
C
O
C
O
HO
O
O
O
O
O
OH
O
C
OH
O
C
C
O
O
O
OH
OCa
O
O
OH
C
O
C
O
OCa
OH
HO
O
O
OCa
OH
OH
O
O
O
O
HO
OCa C HO
C
HO
OCa
O
HO O
O
HO O
ONa
OH
O
O HO
O
HO
OH
C
C
OH
HO
OH
OCa
HO
O
O
ONa
OH
OH
ONa
OH
C
O
O
O
OH
ONa
O
OH
O
OH
OH
OH
O
C
ONa
O
OH
OH
Na-alginate
OH
O
C
ONa
OH
C
O
HO O
O
HO O
OCa
OH
O
OCa
O
O
O
O
HO O
C OH
C
O
O
ONa
OH
O
HO
ONa
O
O
C
C
O
O
C ONa
OH
O
O
O
O
O
C
O
C
OH
OH
O
O
OH O O
O
C
C
O
OH
C
HO
OCa
OH
Outer polymer layer
HO
O
O
O
HO
C
HO
OCa
C
OCa
ONa
HO
O
O
C
HO
OH C
O O
OH
OH
C
O O
O HO
O
ONa
OH
OCa
OH
O
O
OH O
C
O
OH
ONa
OH
O
O
OH
O
OCa
O
ONa
OH
OCa
OH
O
O
OH
O
C
Coating with another polymer
CaCl2
O
Fig. 18 Entrapment of an active ingredient in alginate beads coated with another biopolymer
O
O
Active agent
+
O
O
OH
O
O
HO O
O
C
O
HO O
O
HO O
HO
OH
O
O
C
O OH
O
C
ONa
OH
OCa
OCa
OH
O
HO
C
O
O
O
O
ONa
OH
O
O
196 K. Prasad and J. Kadokawa
Protease
Dye-stained protein
Fig. 19 Release of protein from emulsan–alginate beads
Emulsanalginate bead
+
Emulsanalginate bead
+
+
Dye-stained protein
Free protein
+ Dye
Alginate bead with protein
Protease
Alginate-Based Blends and Nano/Microbeads 197
198
K. Prasad and J. Kadokawa
other biopolymers, konjac glucomannan and chitosan, and their in vitro assessments were done for release of two protein drugs, bovine serum albumin and insulin. It was observed that diffusion of protein was related to the viscosity and swelling properties of konjac glucomannan (Wang and He 2002). Numerical modelling on the release profile of insulin hexamer and amyloglucosidase entrapped in spherical calcium alginate beads was done by Rasmussen et al. (2003). They also modelled the effect of swelling and the size of beads on the protein release rate. Such numerical models are useful for designing desired drug release systems. Xing et al. (2003) evaluated the suitability of coated calcium alginate gel bead entrapped liposome and bee venom peptide as a model drug for colon-specific drug delivery in vitro. They concluded from various studies that the coated calcium alginate gel bead entrapped liposome could be a potential system for colon-specific drug delivery. Apart from the sustained release of drugs as discussed above, the use of alginate beads in slow release of agrochemicals and fertilizers has also been popular in recent years. The release of alachlor (a herbicide from the chloroacetanilide family) from controlled-release formulations based on alginate-montmorillonite bead matrices into aqueous poly(ethylene glycol) solutions and soil at different moisture contents was studied by Nasser et al. (2008). The moisture content of soil, bead diameter, bead weight and the water content were found to affect the release rate. Kenawy and Sakran (1996) used calcium alginate beads for controlled release of 1-naphthalene acetic acid (a plant growth regulator) and of pentachlorophenol (a herbicide) in alkaline and acidic media. They further investigated the role of polyethyleneimine coating of the beads on the release profile. An increased release rate was monitored owing to the polyethyleneimine coating and a slow rate of release from such beads in acidic media was observed. Kulkarni et al. (2000) successfully encapsulated a natural liquid pesticide ‘neem (Azadirachta indica A. Juss) seed oil’ in glutaraldehyde cross-linked sodium alginate beads. They studied the chemical interaction between the ingredient and alginate as well as the effect of cross-linking on the release rate.
3.2.3
Use in Pollution Control
Heavy metal contamination is often found in polluted water and the removal is generally done by various chemical means, including passing through resins and precipitation. There are certain types of biomass present in nature which can form complexes with heavy metals such as Cd(II) through the functional groups present at their surface. Alginic acid and chitosan gel beads have been recognized as one of the most effective adsorbents for eliminating low levels of heavy metal ions from wastewater stream (Gotoh et al. 2004a). Bayramoglu et al. (2002) were successful in achieving the biosorption of Cd(II) by using calcium alginate beads entrapped with a white rot fungus species Lentinus sajor-caju biomass. They reused the biosorbent in three consecutive adsorption-desorption cycles without significant loss of the biosorption capacity. Interestingly, the adsorption of the heavy metal by alginate beads alone was also observed. Later, the same research group revealed the
Alginate-Based Blends and Nano/Microbeads
199
potential use of immobilized microalgae in calcium alginate beads of Chlamydomonas reinhardtii in biosorption of Hg(II), Cd(II) and Pb(II) ions from aqueous solutions using bare calcium alginate beads as a control system (Bayramoglu et al. 2006). Another group attempted to remove Cd(II) ion from aqueous system by using dry biomass of cyanobacteria immobilized in calcium alginate beads (Katircioglu et al. 2007). Owing to the complex formation ability of As(III) with Fe, Zouboulis and Katsoyiannis (2002) modified alginate beads by doping them with hydrous ferric oxides followed by passing As(III)-contaminated water through the bed of the modified beads. Fair removal of As from the contaminated water was observed. Banerjee et al. (2007) used Fe-doped alginate beads for removal of As(III) and As(V). Chen et al. (1997) carried out equilibrium and kinetic studies on Cu ion uptake by calcium alginate beads by varying the pH, ionic strength, initial concentration of Cu ion and the amount of calcium alginate. The ion uptake efficiency was found to increase with increasing pH and decreasing ionic strength. Marine invertebrates and fish are susceptible to nitrate and removal of nitrate from a marine aquarium is essential for the health of the fish and invertebrates. Biological denitrification of nitrate to nitrogen gas in freshwater and a marine aquarium was studied by Tal et al. (2003). The denitrifiers for nitrate removal were successfully immobilized in freeze-dried alginate-starch matrix beads. The ability of protonated alginate beads to take up several heavy metal ions was investigated by Ibanez and Umetsu (2002). The uptake of trivalent Cr, Cu, Zn, Ni and Co was found to be approximately 75, 77, 46, 43 and 35 mg g−1 of beads, respectively. Subsequently, they also studied the uptake of trivalent Cr from aqueous solutions by the protonated dry alginate beads (Ibanez and Umetsu 2004). Calcium alginate beads containing humic acid were found to be efficient for removal of the heavy metals Cd, Cu, Cr, Ni, Mn, Fe and Zn from leachates of industrial waste and tannery effluent (Pandey et al. 2002, Pandey et al. 2003a, Pandey et al. 2003b). Gotoh et al. (2004b) studied the adsorption of Cu and Mn on covalently cross-linked alginate gel beads. Coimmobilization of the freshwater microalga Chlorella vulgaris in alginate beads with the microalgal growth promoting bacterium Azospirillum brasilense under semicontinuous synthetic wastewater culture conditions significantly increased the removal of ammonium and soluble phosphorus ions compared with immobilization of the microalgae alone as observed by Bashan et al. (2002). A methodological approach using mathematical modelling was developed by Veglio et al. (2002) to study adsorption and biosorption of heavy metals. They characterized the behaviour of calcium alginate beads with respect to Cu biosorption. Park et al. (2002) were able to successfully remove H2S using Thiobacillus sp. encapsulated in calcium alginate beads in a continuous H2S-removal bioreactor. Growth of some species of marine microalgae immobilized in calcium alginate beads as well as short term accumulation of heavy metals by them was reported by Moreno-Garrido et al. (2005). Steffan et al. (2005) reported the biodegradation of an azo dye (ethyl orange) by microbial cultures immobilized in alginate beads. Goethite (an Fe-bearing oxide mineral found in soil and other low-temperature environments) is known to bind with Cr(VI). Lazaridis and Charalambous (2005) prepared goethite-conjugated calcium alginate beads to remove Cr(VI) and Cr(III) ions from binary aqueous
200
K. Prasad and J. Kadokawa
solutions. Fiol et al. (2006) also reported the removal of Cr(VI) using 2% grape stalks encapsulated in calcium alginate beads. Calcium alginate prepared from alginic acid having a high b-d-mannuronic acid to a-l-guluronic acid ratio, extracted from Laminaria digitata, was found to adsorb Cu2+, Cd 2+ and Pb2+ in acidic solutions (Papageorgiou et al. 2006). The yeast S. cerevisiae immobilized in calcium alginate beads was used for the separation and subsequent determination of Pt in environmental samples by flow-injection chemiluminescence and electrothermal atomic absorption spectrometry by Godlewska-Zylkiewicz et al. (2008). Organic pollutants are also being treated by alginate beads. p-Chlorophenol and gallic acid are the major organic pollutants present in wastewater. The alginateactivated carbon beads prepared in the presence of CaCl2 were found to adsorb p-chlorophenol selectively from a humic acid solution. Selective adsorption of gallic acid by the beads prepared in the solution of FeCl3 was also observed (Lin et al. 2005). Removal of nitrophenol contaminants using alginate gel beads was reported by Peretz and Cinteza (2007). Niladevi and Prema (2008) carried out the removal of phenol by laccase from Streptomyces psammoticus immobilized on alginate beads. Teerapatsakul et al. (2008) were able to decolorize the dye indigo carmine by laccase entrapped in copper alginate. A selective herbicide, atrazine, was successfully degraded by Rhodococcus erythropolis NI86/21 bacteria encapsulated in alginate beads (Vancov et al. 2005). Polyphenols present in olive mill wastewater were biodegraded with immobilized Candida tropicalis in alginate beads under metabolic induction (Ettayebi et al. 2003). Y. G. Li et al. (2008) found that Pseudomonas delafieldii R-8 immobilized on calcium alginate beads can improve biodesulfurization activity in oil-in-water biphasic systems.
3.2.4
Miscellaneous Use of Alginate Beads Including in Biomedical
Apart from the applications of alginate beads discussed already, research is also going on to explore more application areas for the beads. In this direction, Keshaw et al. (2005) demonstrated the delivery of angiogenic growth factors by incorporating bioactive glass and fibroblasts in alginate beads. Agarose and calcium alginate beads have been used for the crystallization of low molecular mass (14.4 kDa, hen egg-white lysozyme) and high molecular mass (636.0 kDa, alcohol oxidase) proteins by Willaert et al. (2005). Probiotics are living microbial supplements and upon ingestion in certain numbers can exert health benefits. Krasaekoopt et al. (2006) observed survival of probiotics by encapsulating them in chitosan-coated alginate beads in stirred yoghurt from UHT and conventionally treated milk during low-temperature storage. Similarly, Lee et al. (2004) observed survival of the microorganism Lactobacillus bulgaricus KFRI 673 in chitosan-coated calcium alginate microparticles in simulated gastric and intestinal juices. Nisin, a bacteriocin, was immobilized into palmitoylated alginate based Wlms or beads. Wlms or beads incorporating various amounts of nisin could be used efficiently to control the growth of pathogens or microorganisms responsible for spoilage at the surface of round beef or other meat products (Millette et al. 2007). N-(3-Aminopropyl)pyrrole
Alginate-Based Blends and Nano/Microbeads
201
was covalently coupled with alginate in an aqueous-phase reaction by means of a carbodiimide-mediated activation reaction to provide a pyrrole–alginate conjugate for subsequent use in biosensor applications (Fig. 20). The new pyrrole–alginate conjugate was used for the immobilization of polyphenol oxidase onto an electrode surface by physical entrapment resulting from the gellification process and electrochemical polymerization of the pyrrole groups. In addition, biosensors were prepared by entrapment of the polyphenol oxidase in polypyrrole–alginate and regular alginate matrices and their performance for the amperometric determination of catechol chosen as a model analyte was examined by Abu-Rabeah et al. (2005). Similarly, like N-3-aminopropyl pyrrole, biotin (vitamin H or B7) was covalently coupled with alginate in an aqueous-phase reaction by means of a carbodiimidemediated activation reaction to provide a biotin–alginate conjugate for subsequent use in biosensor applications (Fig. 21). The biotin–alginate conjugate was used for the encapsulation of bioluminescent reporter cells into microspheres. A biosensor was prepared by conjugating these biotinylated alginate microspheres to the surface of a streptavidin-coated optical fibre and the performance of the biosensor was demonstrated in the determination of the antibiotic mitomycin C as a model toxin as revealed by Polyak et al. (2004). Desille et al. (2001) fabricated extracorporeal bioartificial liver for pig containing alginate bead entrapped hepatocytes. Calcium alginate beads were utilized to prepare porous keratin sponges by Hamasaki et al. (2007). To prepare such sponges, aqueous solutions of reduced
O
ONa
C O
O
OH
ONa C
OH O
HO
O
OH
O
O
O
OH
C
OH O
O
N
H2N
ONa
O OH
O
O
HO
N-3-amino propyl pyrrole
C ONa
EDAC, NHSS
H N O C
ONa
O
OH
C
HO
N
O
O
O OH
O
HO OH
O
O
O C
OH O ONa
O OH
HO O
O
C ONa
Pyrrole coupled alginate bead EDAC = 1-ethyl-[3-(dimethylamino)propyl]-3-ethylcarbodiimide; NHSS = N-Hydroxysulfosuccinimide
Fig. 20 Synthesis of pyrrole-coupled alginate beads
202
K. Prasad and J. Kadokawa ONa
O C
ONa
O
OH
OH
C
HO
O
HO O
O HO
O O OH
O
O
O OH
O C OH
O EDAC + Sulfo-NHS
ONa
Biotin Hydrazide
C
O
H
ONa
O
HN HN
O
H
S
HN NH
O C
ONa
O OH O HO
O OH
O
C
HO
O
OH
O
O
O C OH
O
ONa
O OH
HO O
O
C
ONa
Biotin coupled alginate bead Fig. 21 Synthesis of biotin-coupled alginate beads
keratin and dried calcium alginate beads were mixed, followed by lyophilization to yield the keratin–calcium alginate complexes. These complexes were further treated with EDTA solution, causing calcium alginate beads to leach out, leaving behind the highly porous keratin sponges. In vitro culture of enzymatically isolated chondrons in sodium alginate beads as a possible model for the initiation of osteoarthritis was reported by Ross et al (2006). Rahman et al. (2005) observed that cotransplantation of encapsulated HepG2 (human hepatocellular liver carcinoma cell) in calcium alginate beads and rat Sertoli cells improves outcome in a thioacetamideinduced rat model of acute hepatic failure. Mishra and Kar (2003) were able to hydrolyse more than 83% of naringin, responsible for the bitter taste of grapefruit juice, by alginate-entrapped naringinase enzyme. Alginate microbeads have recently been introduced in stem cell research. In vitro separation of embryoid-body cells derived from embryonic stem cells into hepatocytes in alginate microbeads was studied by Fang et al. (2007). Ichinose et al. (2005) carried out the detailed examination of cartilage formation and endochondral ossification using human mesenchymal stem cells maintained in alginate beads. The effects of high molecular mass hyaluronan on the distribution and movement of proteoglycan (a type of glycoprotein)
Alginate-Based Blends and Nano/Microbeads
203
formed around rabbit chondrocytes cultured in alginate beads were evaluated and it was found that extrinsic hyaluronan could affect the movement of newly synthesized proteoglycan both in alginate beads and in cartilage tissue (Kikuchi et al. 2001). Sakai et al. (2001) developed a novel type of immunoisolation membrane for a microcapsule-shaped bioartificial pancreas using sol-gel-synthesized silicate. They prepared an aminopropylsilicate membrane derived from 3-aminopropyltrimethoxysilane and tetramethoxysilane on calcium alginate gel beads via electrostatic interaction, where blood islets could be encapsulated and they retain their ability to secrete insulin. Masuda et al. (2003) studied the possibility of the effectiveness of human osteogenic protein-1 in promoting matrix synthesis and matrix formation by rabbit nucleus pulposus and annulus fibrosus cells cultured in alginate beads. The results which they found provided evidence of the ability of human osteogenic protein-1 to stimulate the metabolism of both rabbit annulus fibrosus and nucleus pulposus cells cultured in alginate beads. Grunder et al. (2004) found that bone morphogenetic protein-2 enhances the expression of type II collagen and aggrecan in chondrocytes embedded in alginate beads.
4
Conclusion
Developments made in the field of alginate-based blends and beads in recent years have been summarized in this chapter. The various recent techniques employed for the preparation of the micro- and nano-sized beads, blends and their important applications in different fields were discussed. The application areas covered for the blends and beads were membrane science, slow release of drugs and agrochemicals, biomedicine, process technology including enzyme immobilization, and pollution control. Some advance applications of these systems in research areas such as stem cells and bioartificial organs were also briefly discussed.
References Abu-Rabeah K, Polyak B, Ionescu RE, Cosnier S, Marks RS (2005) Synthesis and characterization of a pyrrole-alginate conjugate and its application in a biosensor construction. Biomacromolecules 6:3313–3318 Albarghouthi M, Fara DA, Saleem M, El-Thaher T, Matalka K, Badwan A (2000) Immobilization of antibodies on alginate-chitosan beads. Int J Pharm 206:23–34 Almeida PF, Almeida AJ (2004) Cross-linked alginate–gelatine beads: a new matrix for controlled release of pindolol. J Control Release 97:431–439 Anal AK, Stevens WF (2005) Chitosan–alginate multilayer beads for controlled release of ampicillin. Int J Pharm 290:45–54 Arica B, Calis S, Kas HS, Sargon MF, Hincal AA (2002) 5-Fluorouracil encapsulated alginate beads for the treatment of breast cancer. Int J Pharm 242:267–269 Ariyakriangkrai W, Tamura H, Tokura S, Rujiravanit R (2005) Preparation and characterization of O-CM chitosan/alginate and N-(carboxyacyl) chitosan/alginate blend fibers. Chitin Chitosan Res 11:180–181
204
K. Prasad and J. Kadokawa
Bajpai SK, Saxena SK, Sharma S (2006) Swelling behavior of barium ions-crosslinked bipolymeric sodium alginate-carboxymethyl guar gum blend beads. React Funct Polym 66:659–666 Banerjee A, Nayak D, Lahiri S (2007) Speciation-dependent studies on removal of arsenic by iron-doped calcium alginate beads. Appl Radiat Isot 65:769–775 Bashan LE, Moreno M, Hernandez JP, Bashan Y (2002) Removal of ammonium and phosphorus ions from synthetic wastewater by the microalgae Chlorella vulgaris coimmobilized in alginate beads with the microalgae growth-promoting bacterium Azospirillum brasilense. Water Res 36:2941–2948 Bayramoglu G, Denizli A, Bektas S, Arica MY (2002) Entrapment of Lentinus sajor-caju into Ca-alginate gel beads for removal of Cd(II) ions from aqueous solution: preparation and biosorption kinetics analysis. Microchem J 72:63–76 Bayramoglu G, Tuzun I, Celik G, Yilmaz M, Arica MY (2006) Biosorption of mercury(II), cadmium(II) and lead(II) ions from aqueous system by microalgae Chlamydomonas reinhardtii immobilized in alginate beads. Int J Miner Process 81:35–43 Bressler E, Pines O, Goldberg I, Braun S (2002) Conversion of fumaric acid to L-malic by sol-gel immobilized saccharomyces cerevisiae in a supported liquid membrane bioreactor. Biotechnol Prog 18:445–450 Carvalho W, Canilha L, Silva SS (2008) Semi-continuous xylose-to-xylitol bioconversion by Ca-alginate entrapped yeast cells in a stirred tank reactor. Bioprocess Biosyst Eng 31:493–498 Castro GR, Kamdar RR, Panilaitis B, Kaplan DL (2005) Triggered release of proteins from emulsan-alginate beads. J Control Release 109:149–157 Caykara T, Demirci S, Eroglu MS, Guven O (2005) Poly(ethylene oxide) and its blends with sodium alginate. Polymer 46:10750–10757 Chen J, Yong F, Yiacoumi S (1997) Equilibrium and kinetic studies of copper ion uptake by calcium alginate. Environ Sci Technol 31:1433–1439 Chen L, Subirade M (2006) Alginate-whey protein granular microspheres as oral delivery vehicles for bioactive compounds. Biomaterials 27:4646–4654 Choi BY, Park HJ, Hwang SJ, Park JB (2002) Preparation of alginate beads for floating drug delivery system: effects of CO2 gas-forming agents. Int J Pharm 239:81–91 Corton E, Piuri M, Battaglini F, Ruzal SM (2000) Characterization of Lactobacillus carbohydrate fermentation activity using immobilized cell technique. Biotechnol Prog 16:59–63 DeGroot AR, Neufeld RJ (2001) Encapsulation of urease in alginate beads and protection from alpha-chymotrypsin with chitosan membranes. Enzyme Microb Technol 29:321–327 Del GP, Colombo P, Colombo G, Russo P, Sonvico F (2005) Mechanisms of formation and disintegration of alginate beads obtained by prilling. Int J Pharm 302:1–9 Desille M, Fremond B, Mahler S, Malledant Y, Seguin P, Bouix A, Lebreton Y, Desbois J, Campion JP, Clement B (2001) Improvement of the neurological status of pigs with acute liver failure by hepatocytes immobilized in alginate gel beads inoculated in an extracorporeal bioartificial liver. Transplant Proc 33:1932–1934 Donati I, Draget KI, Borgogna M, Paoletti S, Skjak-Braek G (2005) Tailor-made alginate bearing galactose moieties on mannuronic residues: selective modification achieved by a chemoenzymatic strategy. Biomacromolecules 6:88–98 Donati I, Haug IJ, Scarpa T, Borgogna M, Draget KI, Skjak-Braek G, Paoletti S (2007) Synergistic effects in semidilute mixed solutions of alginate and lactose-modified chitosan (chitlac) Biomacromolecules 8:957–962 Dong YQ, Zhang L, Shen JN, Song MY, Chen HL (2006a) Preparation of poly(vinyl alcohol)sodium alginate hollow-fiber composite membranes and pervaporation dehydration characterization of aqueous alcohol mixtures. Desalination 193:202–210 Dong ZF, Wang Q, Du YM (2006b) Alginate/gelatin blend films and their properties for drug controlled release. J Membr Sci 280:37–44 Ettayebi K, Errachidi F, Jamai L, Tahri-Jouti MA, Sendide K, Ettayebi M (2003) Biodegradation of polyphenols with immobilized Candida tropicalis under metabolic induction. FEMS Microbiol Lett 223:215–219
Alginate-Based Blends and Nano/Microbeads
205
Fadnavis NW, Sheelu G, Kumar BM, Bhalerao MU, Deshpande AA (2003) Gelatin blends with alginate: gels for lipase immobilization and purification. Biotechnol Prog 19:557–564 Fan LH, Du YM, Huang RH, Wang Q, Wang XH, Zhang LN (2005) Preparation and characterization of alginate/gelatin blend fibers J Appl Polym Sci 96:1625–1629 Fan LH, Du YM, Zhang BZ, Yang JH, Zhou JP, Kennedy JF (2006) Preparation and properties of alginate/carboxymethyl chitosan blend fibers. Carbohydr Polym 65:447–452 Fang S, Qiu Y, Mao L, Shi X, Yu D, Ding Y (2007) Differentiation of embryoid-body cells derived from embryonic stem cells into hepatocytes in alginate microbeads in vitro. Acta Pharmacol Sin 28:1924–1930 Fernandez J, Dhananjeyan MR, Kiwi J, Senuma Y, Hilborn J (2000) Evidence for Fenton photoassisted processes mediated by encapsulated Fe ions at biocompatible pH values. J Phys Chem B 104:5298–5301 Fiol N, Escudero C, Poch J, Villaescusa I (2006) Preliminary studies on Cr(VI) removal from aqueous solution using grape stalk wastes encapsulated in calcium alginate beads in a packed bed up-flow column. React Funct Polym 66:795–807 Fischer FG, Dorfel H (1955) The polyuronic acids of brown algae. Hoppe Seylers Z Physiol Chem 302:186–203 George M, Abraham TE (2006) Polyionic hydrocolloids for the intestinal delivery of protein drugs: Alginate and chitosan. J Control Release 114:1–14 George M, Abraham TE (2007) pH sensitive alginate–guar gum hydrogel for the controlled delivery of protein drugs. Int J Pharm 335:123–129 Gervais TR, Carta G, Gainer JL (2000) Effect of aeration during cell growth on ketone reactions by immobilized yeast. Biotechnol Prog 16:208–212 Ginty PJ, Barry JJ, White LJ, Howdle SM, Shakesheff KM (2008) Controlling protein release from scaffolds using polymer blends and composites. Eur J Pharm Biopharm 68:82–89 Givry S, Prevot V, Duchiron F (2008) Lactic acid production from hemicellulosic hydrolyzate by cells of Lactobacillus bifermentans immobilized in Ca-alginate using response surface methodology. World J Microbiol Biotechnol 24:745–752 Godlewska-Zylkiewicz B, Malejko J, Lesniewska B, Kojlo A (2008) Assessment of immobilized yeast for the separation and determination of platinum in environmental samples by flowinjection chemiluminescence and electrothermal atomic absorption spectrometry. Microchim Acta 163:327–334 Gonzalez-Rodriguez ML, Holgado MA, Sanchez-Lafuente C, Rabasco AM, Fini A (2002) Alginate/chitosan particulate systems for sodium diclofenac release. Int J Pharm 232:225–234 Gotoh T, Matsushima K, Kikuchi K-I (2004a) Preparation of alginate–chitosan hybrid gel beads and adsorption of divalent metal ions. Chemosphere 55:135–140 Gotoh T, Matsushima K, Kikuchi K-I (2004b) Preparation of alginate-chitosan hybrid gel beads and adsorption of divalent metal ions. Chemosphere 55:57–64 Grant GT, Morris ER, Rees DA, Smith PJC, Thom D (1973) Biological interactions between polysaccharides and divalent cations: the egg- box model. FEBS Lett 32:195–198 Griffin DR, Yang F, Carta G, Gainer JL (1998) Asymmetric reduction of acetophenone with Calcium-alginate-entrapped baker’s yeast in organic solvents. Biotechnol Prog 14:588–593 Grunder T, Gaissmaier C, Fritz J, Stoop R, Hortschansky P, Mollenhauer J, Aicher WK (2004) Bone morphogenetic protein (BMP)-2 enhances the expression of type II collagen and aggrecan in chondrocytes embedded in alginate beads. Osteoarthritis Cartil 12:559–567 Gu F, Amsden B, Neufeld R (2004) Sustained delivery of vascular endothelial growth factor with alginate beads. J Control Release 96:463–472 Hamasaki S, Tachibana A, Tada D, Yamauchi K, Tanabe T (2007) Fabrication of highly porous keratin sponges by freeze-drying in the presence of calcium alginate beads. Mater Sci Eng C 28:1250–1254 Hann EC, Sigmund AE, Hennessey SM, Gavagan JE, Short DR, Ben-Bassat A, Chauhan S, Fallon RD, Payne MS, DiCosimo R (2002) Optimization of an immobilized-cell bio-catalyst for production of 4-cyanopentanoic acid. Org Process Res Dev 6:492–496
206
K. Prasad and J. Kadokawa
Haug A, Larsen B, Smidsrod, O (1967) Studies on the Sequence of uronic acid residues in alginic acid. Acta Chem Scand 21:691–704 Hsieh H-J, Tung K-Y, Nair GR, Chu I-M, Wu W-T (2007) Production of ascorbic acid glucoside by alginate-entrapped mycelia of Aspergillus niger. Appl Microbiol Biotechnol 77:53–60 Hsu SH, Whu SW, Hsieh SC, Tsai CL, Chen DC, Tan TS (2004) Evaluation of chitosan-alginatehyaluronate complexes modified by an RGD-containing protein as tissue-engineering scaffolds for cartilage regeneration. Artif Organs 28:693–703 Ibanez JP, Umetsu Y (2002) Potential of protonated alginate beads for heavy metals uptake. Hydrometallurgy 64:89–99 Ibanez JP, Umetsu Y (2004) Uptake of trivalent chromium from aqueous solutions using protonated dry alginate beads. Hydrometallurgy 72:327–334 Ichinose S, Yamagata K, Sekiya I, Muneta T, Tagami M (2005) Detailed examination of cartilage formation and endochondral ossification using human mesenchymal stem cells. Clin Exp Pharmacol Physiol 32:561–570 Iizuka Y, Kato R, Shibasaki-Kitakawa N, Yonemoto T (2005) Combination of extractive solvent addition and immobilization culture for continuous production of scopoletin by tobacco cells. Biotechnol Prog 21:603–607 Iskakov RM, Kikuchi A, Okano T (2002) Time-programmed pulsatile release of dextran from calciumalginate gel beads coated with carboxy-n-propylacrylamide copolymers. J Control Release 80:57–68 Iwasaki N, Yamane ST, Majima T, Kasahara Y, Minami A, Harada K, Nonaka Sachiko, Maekawa N, Tamura H, Tokura S, Shiono M, Monde K, Nishimura S-I (2004) Feasibility of polysaccharide hybrid materials for scaffolds in cartilage tissue engineering: Evaluation of chondrocyte adhesion to polyion complex fibers prepared from alginate and chitosan. Biomacromolecules 5:828–833 Iwata H, Simada H, Fukuma E, Ibii T, Sato H (2004) Bioartificial pancreas research in Japan. Artif Organs 28:45–52 Kadokawa J, Saitou S, Shoda S (2005) Preparation of alginate-polymethacrylate hybrid material by radical polymerisation of cationic methacrylate monomer in the presence of sodium alginate. Carbohydr Polym 60:253–258 Kalyani S, Smitha B, Sridhar S, Krishnaiah A (2006) Blend membranes of sodium alginate and hydroxyethylcellulose for pervaporation-based enrichment of t-butyl alcohol. Carbohydr Polym 64:425–432 Kanti P, Srigowri K, Madhuri J, Smitha B, Sridhar S (2004) Dehydration of ethanol through blend membranes of chitosan and sodium alginate by pervaporation. Sep Purif Technol 40:259–266 Katircioglu H, Aslim B, Turker AR, Atici T, Beyatli Y (2007) Removal of cadmium(II) ion from aqueous system by dry biomass, immobilized live and heat-inactivated Oscillatoria sp. H1 isolated from freshwater (Mogan Lake). Bioresour Technol. doi:10.1016/j.biortech.2007.08.068 Kaul P, Banerjee A, Banerjee UC (2006) Stereoselective nitrile hydrolysis by immobilized wholecell bio-catalyst. Biomacromolecules 7:1536–1541 Kenawy E-R, Sakran MA (1996) Controlled release formulations of agrochemicals from calcium alginate. Ind Eng Chem Res 35:3726–3729 Keshaw H, Forbes A, Day RM (2005) Release of angiogenic growth factors from cells encapsulated in alginate beads with bioactive glass. Biomaterials 26:4171–4179 Kikuchi T, Yamada H, Fujikawa K (2001) Effects of high molecular weight hyaluronan on the distribution and movement of proteoglycan around chondrocytes cultured in alginate beads. Osteoarthritis Cartil 9:351–356 Kim MH, Kim JC Lee HY, Kim JD, Yang JH (2005) Release property of temperature-sensitive alginate beads containing poly (N-isopropylacrylamide). Colloids Surf B 46:57–61 Kim YS, Kim HW, Lee SH, Shin KS, Hur HW, Rhee YH (2007) Preparation of alginate–quaternary ammonium complex beads and evaluation of their antimicrobial activity. Int J Biol Macromol 41:36–41 Krasaekoopt W, Bhandari B, Deeth HC (2006) Survival of probiotics encapsulated in chitosancoated alginate beads in yoghurt from UHT- and conventionally treated milk during storage. LWT Food Sci Technol 39:177–183
Alginate-Based Blends and Nano/Microbeads
207
Kulkarni AR, Soppimath KS, Aminabhavi TM, Dave AM, Mehta MH (2000) Glutaraldehyde crosslinked sodium alginate beads containing liquid pesticide for soil application. J Control Release 63:97–105 Kulkarni AR, Soppimath KS, Aminabhavi TM, Rudzinski WE (2001) In-vitro release kinetics of cefadroxil-loaded sodium alginate interpenetrating network beads. Eur J Pharm Biopharm 51:127–133 Lazaridis NK, Charalambous C (2005) Sorptive removal of trivalent and hexavalent chromium from binary aqueous solutions by composite alginate–goethite beads. Water Res 39:4385–4396 Lee JS, Cha DS, Park HJ (2004) Survival of freeze-dried lactobacillus bulgaricus KFRI 673 in chitosan-coated Calcium alginate microparticles. J Agric Food Chem 52:7300–7305 Li B, Zhu JB, Zheng CL, Li WGB, Zhu JB, Zheng CL, Gong W (2008) A novel system for threepulse drug release based on “tablets in capsule” device. Int J Pharm 352:159–164 Li XY, Jin LJ, McAllister TA, Stanford K, Xu JY, Lu YN, Zhen YH, Sun YX, Xu YP (2007) Chitosan-alginate microcapsules for oral delivery of egg yolk immunoglobulin (IgY). J Agric Food Chem 55:2911–2917 Li YG, Xing JM, Xiong XC, Li WL, Gao HS, Liu HZ (2008) Improvement of biodesulfurization activity of alginate immobilized cells in biphasic systems. J Ind Microbiol Biotechnol 35:145–150 Lin Y-B, Fugetsu B, Terui N, Tanaka S (2005) Removal of organic compounds by alginate gel beads with entrapped activated carbon. J Hazard Mater B 120:237–241 Llaneras M (2000) Brown algae of economic significance. In: Martin RE, Carter EP, Flick GJ, Davis LM (eds) Marine and fresh water product hand book. CRC, Boca Raton, pp 531–541 Martin AR, Shonnard D, Pannuri S, Kamat S (2007) Characterization of free and immobilized (S)-aminotransferase for acetophenone production. Appl Microbiol Biotechnol 76:843–851 Masuda K, Takegami K, An H, Kumano F, Chiba K, Andersson GBJ, Schmid T, Thonar E (2003) Recombinant osteogenic protein-1 upregulates extracellular matrix metabolism by rabbit annulus fibrosus and nucleus pulposus cells cultured in alginate beads. J Orthop Res 21:922–930 Meena R, Chhatbar M, Prasad K, Siddhanta AK (2008) Development of a robust hydrogel system based on agar and sodium alginate blend. Polym Int 57:329–336 Mi F-L, Sung H-W, Shyu S-S (2002) Drug release from chitosan–alginate complex beads reinforced by a naturally occurring cross-linking agent. Carbohydr Polym 48:61–72 Milagre HMS, Milagre CDF, Paulo J. S. Moran PJS, Santana MHA, Rodrigues JAR (2006) Asymmetric bio-reduction of ethyl 3-halo-2-oxo-4-phenylbutanoate by Saccharomyces cerevisiae immobilized in Ca-alginate beads with double gel layer. Org Process Res Dev 10:611–617 Millette M, Tien CL, Smoragiewicz W, Lacroix M (2007) Inhibition of Staphylococcus aureus on beef by nisin-containing modiWed alginate Wlms and beads. Food Control 18:878–884 Mishra P, Kar R (2003) Treatment of grapefruit juice for bitterness removal by Amberlite IR 120 and Amberlite IR 400 and alginate entrapped naringinase enzyme. J Food Sci 68:1229–1233 Mofidi N, Aghai-Moghadam M, Sarbolouki MN (2000) Mass preparation and characterization of alginate microspheres. Process Biochem 35:885–888 Morch YA, Donati I, Strand BL, Skjak-Braek G (2007) Molecular engineering as an approach to design new functional properties of alginate. Biomacromolecules 8:2809–2814 Moreno-Garrido I, Campana O, Lubian LM, Blasco J (2005) Calcium alginate immobilized marine microalgae: Experiments on growth and short-term heavy metal accumulation. Mar Pollut Bull 51:823–829 Mohammad BT, Bustard MT (2008) Fed batch bioconversion of 2-propanol by a solvent tolerant strain of Alcaligenes faecalis entrapped in Ca-alginate gel. J Ind Microbiol Biotechnol 35:677–684 Murata Y, Sasaki N, Miyamoto E, Kawashima S (2000) Use of floating alginate gel beads for stomach-specific drug delivery. Eur J Pharm Biopharm 50:221–226 Murata Y, Kontani Y, Ohmae H, Kawashima S (2002) Behavior of alginate gel beads containing chitosan salt prepared with water-soluble vitamins. Eur J Pharm Biopharm 53:249–251 Naganagouda K, Mulimani VH (2006) Gelatin blends with alginate: gel fibers for a-galactosidase immobilization and its application in reduction of non-digestible oligosaccharides in soymilk. Process Biochem 41:1903–1907
208
K. Prasad and J. Kadokawa
Naidu BVK, Aminabhavi TM (2005) Pervaporation separation of water/2-propanol mixtures by use of the blend membranes of sodium alginate and (hydroxyethyl) cellulose. Ind Eng Chem Res 44:7481–7489 Naidu BVK, Malladi S, Raju KVSN, Aminabhavi TM (2005a) Thermal, viscoelastic, solution and membrane properties of sodium alginate/hydroxyethylcellulose blends. Carbohydr Polym 61:52–60 Naidu BVK, Rao K.S.V, Aminabhavi TM (2005b) Pervaporation separation of water + 1,4-dioxane and water + tetrahydrofuran mixtures using sodium alginate and its blend membranes with hydroxyethylcellulose – a comparative study. J Membr Sci 260:131–141 Nedovic VA, Obradovic B, Ukalovic IL, Trifunovic O, Pesic R, Bugarski B (2001) Electrostatic generation of alginate microbeads loaded with brewing yeast. Process Biochem 37:17–22 Nasser A, Mingelgrin U, Gerstl Z (2008) Effect of soil moisture on the release of alachlor from alginate-based controlled-release formulations. J Agric Food Chem 56:1322–1327 Niladevi KN, Prema P (2008) Immobilization of laccase from Streptomyces psammoticus and its application in phenol removal using packed bed reactor. World J Microbiol Biotechnol 24(7):1215–1222 Pandey AK, Pandey SD, Misra V (2002) Removal of toxic metals from leachates from hazardous solid wastes and reduction of toxicity to microtox by the use of calcium alginate beads containing humic acid. Ecotoxicol Environ Saf 52:92–96 Pandey AK, Pandey SD, Misra V, Devi S (2003a) Role of humic acid entrapped calcium alginate beads in removal of heavy metals. J Hazard Mater 98:177–181 Pandey AK., Pandey SD, Misra V, Srimal AK (2003b) Removal of chromium and reduction of toxicity to microtox system from tannery effluent by the use of calcium alginate beads containing humic acid. Chemosphere 51:329–333 Papageorgiou SK, Katsaros FK, Kouvelos EP, Nolan JW, Deit HL, Kanellopoulos NK (2006) Heavy metal sorption by calcium alginate beads from Laminaria digitata. J Hazard Mater B 137:1765–1772 Park D-H, Cha J-M, Ryu H-Won, Lee G-W, Yu E-Y, Rheea J-I, Park J-J, Kimd S-W, Lee I-W, Joe Y-I, Ryu Y-W, Hurg B-K, Park J-K, Park K (2002) Hydrogen sulfide removal utilizing immobilized Thiobacillus sp. IW with Ca-alginate bead. Biochem Eng J 11:167–173 Pasparakis G, Bouropoulos N (2006) Swelling studies and in vitro release of verapamil from calcium alginate and calcium alginate–chitosan beads. Int J Pharm 323:34–42 Peretz S, Cinteza O (2007) Removal of some nitrophenol contaminants using alginate gel beads. Colloids Surf A 319:165–172 Phisalaphong M, Suwanmajo T, Tammarate P (2008) Synthesis and characterization of bacterial cellulose/alginate blend membranes. J Appl Polym Sci 107:3419–3424 Pongjanyakul T, Puttipipatkhachorn S (2007) Xanthan–alginate composite gel beads: molecular interaction and in vitro characterization. Int J Pharm 331:61–71 Pongjanyakul T, Sungthongjeen S, Puttipipatkhachorn S (2006) Modulation of drug release from glyceryl palmitostearate–alginate beads via heat treatment. Int J Pharm 319:20–28 Polyak B, Geresh S, Marks RS (2004) Synthesis and characterization of a biotin-alginate conjugate and its application in a biosensor construction. Biomacromolecules 5:389–396 Rahman TM, Diakanov I, Selden C, Hodgson H (2005) Co-transplantation of encapsulated HepG2 and rat Sertoli cells improves outcome in a thioacetamide induced rat model of acute hepatic failure. Transplant Int 18:1001–1009 Ramesh Babu V, Sairam M, Hosamani KM, Aminabhavi TM (2007) Preparation of sodium alginate– methylcellulose blend microspheres for controlled release of nifedipine. Carbohydr Polym 69:241–250 Rinaudo M (2008) Main properties and current applications of some polysaccharides as biomaterials. Polym Int 57:397–430 Rokstad AM, Donati I, Borgogna M, Oberholzer J, Strand BL, Espevik T, Skjak-Braek G (2006) Cell-compatible covalently reinforced beads obtained from a chemoenzymatically engineered alginate. Biomaterials 27:4726–4737
Alginate-Based Blends and Nano/Microbeads
209
Ross JM Sherwin AF, Poole CA (2006) In vitro culture of enzymatically isolated chondrons: a possible model for the initiation of osteoarthritis, J Anat 209:793–806 Rousseau I, Cerf D L, Picton L, Argillier JF, Muller G (2004) Entrapment and release of sodium polystyrene sulfonate (SPS) from calcium alginate gel beads. Eur Polym J 40:2709–2715 Roy I, Gupta MN (2004) Hydrolysis of starch by a mixture of glucoamylase and pullulanase entrapped individually in calcium alginate beads. Enzyme Microb Technol 34:26–32 Roy I, Sardar M, Gupta MN (2005) Cross-linked alginate–guar gum beads as fluidized bed affinity media for purification of jacalin. Biochem Eng J 23:193–198 Ryu S-A, Kim CS, Kim H-J, Baek DH, Oh D-K (2003) Continuous D-tagatose production by immobilized thermostable L-arabinose isomerase in a packed-bed bioreactor. Biotechnol Prog 19:1643–1647 Sakai S, Ono T, Ijima H, Kawakami K (2001) Synthesis and transport characterization of alginate/ aminopropylsilicate/alginate microcapsule: application to bioartificial pancreas. Biomaterials 22:2827–2834 Sakai S, Ono T, Ijima H, Kawakami K (2002) Alginate/aminopropyl-silicate/alginate membrane immunoisolatability and insulin secretion of encapsulated islets. Biotechnol Prog 18:401–403 Sakai S, Kawakami K (2007) Synthesis and characterization of both ionically and enzymatically cross-linkable alginate. Acta Biomater 3:495–501 Sanli O, Ay N, Isiklan N (2007) Release characteristics of diclofenac sodium from poly(vinyl alcohol)/sodium alginate and poly(vinyl alcohol)-grafted-poly(acrylamide)/sodium alginate blend beads. Eur J Pharm Biopharm 65:204–214 Sarmento B, Ribeiro A, Veiga F, Sampaio P, Neufeld R, Ferreira D (2007) Alginate/chitosan nanoparticles are effective for oral insulin delivery. Pharm Res 24:2198–2206 Seifert DB, Phillips JA (1997) Production of small, monodispersed alginate beads for cell immobilization. Biotechnol Prog 13:562–568 Shu XZ, Zhu KJ (2002) The release behavior of brilliant blue from calcium–alginate gel beads coated by chitosan: the preparation method effect. Eur J Pharm Biopharm 53:193–201 Smitha B, Sridhar S, Khan AA (2005) Chitosan–sodium alginate polyion complexes as fuel cell membranes. Eur Polym J 41:1859–1866 Rasmussen MR, Snabe T, Pedersen LH (2003) Numerical modelling of insulin and amyloglucosidase release from swelling Ca’alginate beads J Control Release 91:395–405 Solak EK, Asman G, Camurlu P, Sanli O (2008) Sorption, diffusion, and pervaporation characteristics of dimethylformamide/water mixtures using sodium alginate/polyvinyl pyrrolidone blend membranes. Vacuum 82:579–587 Steffan S, Bardi L, Marzon M (2005) Azo dye biodegradation by microbial cultures immobilized in alginate beads. Environ Int 31:201– 205 Sugiura S, Oda T, Izumida Y, Aoyagi Y, Satake M, Ochiai A, Ohkohchi N, Nakajima M (2005) Size control of calcium alginate beads containing living cells using micro-nozzle array Biomaterials 26:3327–3331 Sussman G (2006) Management of the wound environment with dressings and topical agents. In: Sussman C, Bates-Jensen BM (eds) Wound care: a collaborative practice manual. Lippincott Williams & Wilkins, Philadelphia, pp 250–267 Tal Y, Nussinovitch A, van Rijn J (2003) Nitrate removal in aquariums by immobilized Pseudomonas. Biotechnol Prog 19:1019–1021 Teerapatsakul C, Bucke C, Parra R, Keshavarz T, Chitradon L (2008) Dye decolorisation by laccase entrapped in copper alginate. World J Microbiol Biotechnol 24:1367–1374 Ueng WN, Yuan L-J, Lee N, Lin S-S, Chan E-C, Weng J-H (2004) In vivo study of biodegradable alginate antibiotic beads in rabbits Steve. J Orthop Res 22:592–599 Um E, Lee D-S, Pyo H-B, Park J-K (2008) Continuous generation of hydrogel beads and encapsulation of biological materials using a microfluidic droplet-merging channel. Microfluid Nanofluid 5:541–549 Vancov T, Jury K, Van Zwieten L (2005) Atrazine degradation by encapsulated Rhodococcus erythropolis NI86/21. J Appl Microbiol 99:767–775
210
K. Prasad and J. Kadokawa
Veglio F, Esposito A, Reverberi AP (2002) Copper adsorption on calcium alginate beads: equilibrium pH-related models. Hydrometallurgy 65:43–57 Wang K, He Z (2002) Alginate/konjac glucomannan/chitosan beads as controlled release matrix. Int J Pharm 244:117–126 Wang Q, Du YM, Hu XW, Yang JH, Fan LH, Feng T (2006) Preparation of alginate/soy protein isolate blend fibers through a novel coagulating bath. J Appl Polym Sci 101:425–431 Willaert R, Zegers I, Wyns L, Sleutel M (2005) Protein crystallization in hydrogel beads. Acta Crystallogr D 61:1280–1288 Won K, Kima S, Kim K-J, Park HW, Moon S-J (2005) Optimization of lipase entrapment in Ca-alginate gel beads. Process Biochem 40:2149–2154 Wong TW, Chan LW, Kho SB, Heng PWS (2002) Design of controlled-release solid dosage forms of alginate and chitosan using microwave. J Control Release 84:99–114 Wu S-Y, Lin C-N, Chang J-S, Lee K-S, Lin P-J (2002) Microbial hydrogen production with immobilized sewage sludge. Biotechnol Prog 18:921–926 Xing L, Dawei C, Liping X, Rongqing Z (2003) Oral colon-specific drug delivery for bee venom peptide: development of a coated calcium alginate gel beads-entrapped liposome. J Control Release 93:293–300 Xu JB, Bartley JP, Johnson RA (2003) Preparation and characterization of alginate–carrageenan hydrogel films crosslinked using a water-soluble carbodiimide (WSC). J Membr Sci 218:131–146 Xu SW, Jiang ZY, Lu Y, Wu H, Yuan WK, (2006) Preparation and catalytic properties of novel alginate-silica-dehydrogenase hybrid biocomposite beads. Ind Eng Chem Res 45:511–517 Xu Y, Zhan C, Fan L, Wang L, Zheng H (2007) Preparation of dual crosslinked alginate-chitosan blend gel beads and in vitro controlled release in oral site-specific drug delivery system. Int J Pharm 336:329–337 Xulu PM, Filipcsei G, Zrinyi M (2000) Preparation and responsive properties of magnetically soft poly(N-isopropylacrylamide) gels. Macromolecules 33:1716–1719 Yadav GD, Jadhav SR (2005) Synthesis of reusable lipases by immobilization on hexagonal mesoporous silica and encapsulation in calcium alginate: trans esterification in non-aqueous medium. Microporous Mesoporous Mater 86:215–222 Yang G, Zhang L, Peng T, Zhong W (2000) Effects of Ca2C bridge cross-linking on structure and pervaporation of cellulose/alginate blend membranes. J Membr Sci 175:53–60 Zhao Y, Carvajal MT, Won Y-Y, Harris MT (2007) Preparation of calcium alginate microgel beads in an electrodispersion reactor using an internal source of calcium carbonate nanoparticles. Langmuir 23:12489–12496 Zohar-Perez C, Chet I, Nussinovitch A (2004) Irregular textural features of dried alginate–filler beads. Food Hydrocolloids 18:249–258 Zouboulis AI, Katsoyiannis IA (2002) Arsenic removal using iron oxide loaded alginate beads. Ind Eng Chem Res 41:6149–6155
Applications of Alginates in Food Iain A. Brownlee, Chris J. Seal, Matthew Wilcox, Peter W. Dettmar, and Jeff P. Pearson
Contents 1 2 3 4 5 6
Introduction........................................................................................................................ Gelling Agents ................................................................................................................... Thickening, Stabilising and Emulsifying Agents .............................................................. Encapsulation and Immobilisation..................................................................................... Food Coating...................................................................................................................... Physiological Effects of Dietary Alginates ........................................................................ 6.1 Alginate Safety ......................................................................................................... 6.2 Effect on Colonic Microflora .................................................................................... 6.3 Reduction of Toxicity of Colonic Luminal Contents ................................................ 6.4 Effect of Alginate on Satiety..................................................................................... 6.5 Effect on Intestinal Absorption Rates ....................................................................... 6.6 Effect on Plasma Cholesterol .................................................................................... 6.7 Effect on Glycaemic and Insulinaemic Responses ................................................... 7 Summary ............................................................................................................................ References ................................................................................................................................
212 213 214 215 217 218 218 218 219 220 220 221 222 222 223
Abstract Alginates are a group of viscous polysaccharides derived from brown seaweeds and produced as an extracellular matrix by some bacterial species. Traditionally, alginates have been used as thickeners, emulsifiers and stabilising agents in a number of food applications. A number of new food applications are emerging for alginates, based on their unique biochemical and biophysical properties.
I.A. Brownlee () and C.J. Seal School of Agriculture, Food & Rural Development, Newcastle University, Newcastle upon Tyne NE1 7RU, UK e-mail: [email protected] M. Wilcox and J.P. Pearson Institute of Cell & Molecular Biosciences, Medical School, Newcastle University, Newcastle upon Tyne NE2 4HH, UK P.W. Dettmar Technostics Ltd, The Deep Business Centre, Kingston upon Hull HU1 4BG, UK
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_9, © Springer-Verlag Berlin Heidelberg 2009
211
212
I.A. Brownlee et al.
These applications are discussed within this review, as are the nutritional properties of dietary alginates, and their physiological actions in humans.
1
Introduction
Alginate is a polyuronan that is isolated from the cell walls of a number of brown algal and bacterial species around the world, and is also produced as an extracellular matrix by certain bacteria (Stokke et al. 2000). Alginates are used in a number of large-scale industrial processes (e.g. as an ingredient in shoe polish and as an important factor in the dye industry (Jensen 1993; Yamamoto et al. 1992) and have medical uses, such as cell microencapsulation (Uludag et al. 2000), as microsphere vectors for drug delivery (Skaugrud et al. 1999), for making dental impressions (Ertesvag and Valla 1998), and as the active ingredient in absorbent dressings (Ingram et al. 1998) and antireflux therapies (Bryan et al. 2001; Mandel et al. 2000). However, more than 6 times the amount of income from these applications is generated worldwide by the use of seaweed and seaweed products in the food industry (Jensen 1993). While large quantities of seaweed are eaten in eastern Asia, the inclusion of algal phycocolloids, (e.g. alginate, carrageenan and agar) as additives is how they are commonly consumed in the Western diet. (Ertesvag and Valla 1998; Moe et al. 1995). As food additives that have been demonstrated as safe and approved for use throughout the EU, alginates are given E numbers (Food Standards Agency 2002). Alginic acid is E400, while the sodium, potassium, ammonium and calcium salts are E401–E404, respectively (Smith and LilyHong 2003). The most widely used of these in the food industry is the sodium salt (Mancini et al. 2002). Esters of alginic acids, propylene glycol alginates (PGAs), are produced industrially with various degrees of esterification and viscosity, and are used for food uses similar to those of other forms of alginate, but at a lower concentration (Baeza et al. 2004). They are identified with the number E405 (Smith and LilyHong 2003). Food industry applications of alginates are entirely dependent on the biophysical properties of the alginates. These properties are derived from the unique chemical structure of alginates. Alginates consist of 1→4-linked a-l-guluronic acid (G) and b-d-mannuronic acid (M) pyranose residues in an unbranched chain. These residues can combine to form G-rich (G blocks) or MG areas (MG blocks), as well as M-rich areas (M blocks), which are especially predominant in bacterial alginates. G blocks are believed to be important to alginate structure as a result of their Ca2+/H+ binding capability, which allows alginates to form gels in the presence of Ca2+ or H+ ions at low temperatures. MG blocks allow polysaccharide chain flexibility and therefore tend to reduce alginate solution viscosity (Smidsrod and Draget 1996). These unique gelling properties of alginates are vital to many of their applications in the food industry. Table 1 compares a number of polysaccharides commonly used in thickening applications and their biophysical properties. The major uses of alginates in the food industry are discussed in the following sections, along with potential applications as suggested by food science literature. Finally, the effects that dietary alginates have on human physiological processes are discussed in Sect. 6.
Applications of Alginates in Food
213
Table 1 Comparison of the biophysical and biochemical properties of a range of viscous polysaccharides used by the food industry. (Data collected from Smith and LilyHong 2003; De Ruiter and Rudolph 1997; de Vries 2004; Sworn 2004; Ward 2004
Alginate
Source
Constituent sugars Properties in gel or solution
Brown seaweed
Mannuronic acid Guluronic acid
Propylene glycol Brown seaweed alginate Pectin Citrus fruit, apples, sugar beets, sunflowers Carrageenan
Red seaweeds
Guar gum
Cyamposis tetragonolobus seeds
Gum arabic
Acacia sap
2
Mannuronic acid Guluronic acid Galacturonic acid Rhamnose Galactose Arabinose Sulphated d-galactose l-Anhydrogalactose
Forms heat-stable gels rapidly at low temperature without prior heating. Gels at low pH or high Ca2+ concentration Does not gel. Useful in stabilising foams Acid-stable gel. Requires acid/sugar/ calcium presence for gelling
Form gels at low concentrations compared with other gelling agents Easily soluble in sugar solutions. Gels are degraded at low pH and high-temperature conditions l-Mannose Hydrates rapidly in cold water, forming highly viscous aqueous d-Galactose solutions. Gels in the presence of other gums (e.g. locust bean gum, xanthan) d-Galactose Has extremely high solubility (up to d-Glucuronic acid 50% aqueous solutions) with low l-Rhamnose viscosity in cold water l-Arabinose
Gelling Agents
Restructuring of foods is commonly used by the food industry as a means of using less aesthetically appealing produce (e.g. cuts of meat with high fat or connective tissue content, or misshapen or missized fruits). By increasing the aesthetic quality of foods, restructured forms can often offer novel products that help meet customer demand. Restructured foods can be produced in any shape or size, which allows production of foodstuffs that are of a more uniform or attractive aesthetic nature. At the same time, the texture and structural properties of the foods can be controlled to produce more desirable products (Ranken 2000). Postharvest losses of fruits and vegetables can be high, so development of processes that can produce value-added, shelf-stable products that can be sold throughout the year is desirable (Mancini and McHugh 2000). Alginates are available at a wide range of viscosities, offer stability to foodstuffs under both high and low temperatures, and therefore have a wide range of uses as gelling agents. Alginate gelation rates and gel strength can be
214
I.A. Brownlee et al.
controlled by the concentration of Ca2+ or H+ in solution. Alginate gel strength is also governed by the number of G blocks within the polyuronate chain. As alginates form gels at low temperatures, this is particularly useful in the restructuring of foodstuffs that may become damaged or oxidised under high temperatures (e.g. meat products, fruits and vegetables). The most common restructured foods produced using alginates are reconstituted onion rings and pimento sections for use in olives (Brownlee et al. 2005). For both of these products, alginates allow the production of products of uniform size and consistency. In these two cases, restructuring also greatly facilitates mass production of these products. Alginates also have a number of similar applications in meat (Ranken 2000;CongGui et al. 2006; Devatkal and Mendiratta 2001; Boles and Shand 1999), seafood (Suklim et al. 2004), fruit (Mancini and McHugh 2000; Das Gupta et al. 2007; Grizotto et al. 2006, 2007), vegetable (Manjunatha and Gupta 2006) and some extruded food products (e.g. pastas and noodles (Brownlee et al. 2005).
3
Thickening, Stabilising and Emulsifying Agents
Alginates are commonly used as thickening and agents in jams, marmalades and fruit sauces, as alginate–pectin interactions are heat-reversible and give a higher viscosity than does either individual component (Brownlee et al. 2005). Alginates are also used to thicken desserts and savoury sauces, including mayonnaise (Mancini et al. 2002; Wendin et al. 1997; Gujral et al. 2001). PGAs are often used for this application, but at concentrations lower than those of standard alginate salts. This is due to their high stability and relatively low viscosity. Use of alginates on their own or in conjunction with other thickening agents has been shown to improve the acceptability of a number of low-fat processed foods (Mancini et al. 2002; Wendin et al. 1997; Kumar and Sahoo 2006; Kumar et al. 2007; Lin and Keeton 1998). Hydrocolloid use in this manner aids retention of moisture and improves food texture. This can result in an improvement to the organoleptic qualities of the food products, thus improving consumer acceptance. Alginates form stable gels at high and low temperatures and at low pH (Smidsrod and Draget 1996). As a result, they can be used for a number of stability applications in food processing (Paraskevopoulou et al. 2006, 2005; Gomez-Diaz and Navaza 2004; Ferreira et al. 2005; Hubbermann et al. 2006). Routine use of alginates in bakery creams endows the cream with freeze/thaw stability and reduces separation of the solid and liquid components (syneresis) (Brownlee et al. 2005). Alginates are used in combination with other hydrocolloids to thicken and stabilise ice cream. While this allows control of the product’s viscosity, it also increases heat-shock resistance, reduces shrinkage and ice crystal formation, and endows the ice cream with the desired melting characteristics (Regand and Goff 2003). PGAs are commonly used to maintain foam stability, including applications in mousse and other desserts (Baeza et al. 2004; Sarker and Wilde 1999). The largest
Applications of Alginates in Food
215
food use of PGA is within the brewing industry, where PGA added to different beers and lagers stabilises the froth head when poured, while also protecting it from foam-negative contaminants (Ferreira et al. 2005). By definition, hydrocolloids are water-soluble and very hydrophilic. However, previous studies have suggested that, within their use as stabilisers in food oil–water emulsions, they exhibit the ability to precipitate/adsorb onto oil droplets and sterically stabilise emulsions against flocculation and coalescence (Huang et al. 2001). PGAs have previously been demonstrated to be better emulsifiers than methylcellulose compounds and locust bean and guar gums, and to be comparable with other commonly used plant-derived polysaccharides (i.e. pectin and gum arabic) on a weight-for-weight basis (Huang et al. 2001; Gaonkar 1991). These factors are useful in simple food emulsions in which PGA and alginates are utilised, such as mayonnaise and other dressings (Mancini et al. 2002; Wendin et al. 1997; Paraskevopoulou et al. 2005).
4
Encapsulation and Immobilisation
One of the more recent uses of alginates is in encapsulation technologies. Outside the food industry, these technologies have been utilised in a wide range of drug delivery applications (George and Abraham 2006; Lee et al. 2003). Microencapsulation of insulin-producing cells for injection into type I diabetics has also been shown to be an effective treatment (de Vos et al. 2006). For this role, alginate is used to immunoisolate injected cells, thereby acting as a physical barrier against host immune sampling and response. This technology, therefore, has potential for a range of other uses in cell transplantation therapies (Uludag et al. 2000). Within the food industry, alginate encapsulation and immobilisation technologies are used for a variety of purposes, including food processing, food functionality and product acceptability. These immobilisation technologies can enhance productivity as a result of continuous operation and reuse of the entrapped cells or enzymes (Groboillot et al. 1994). Immobilisation or encapsulation technology is used to produce a wide range of bacterial metabolites, including enzymes, amino acids, organic acids (e.g. acetic acid) and alcohols (Norton and Vuillemard 1994). Alginates are commonly used as an immobilisation medium for whole cell or isolated enzyme preparations in industrial processes requiring enzymatic activity. Immobilisation in this manner is a simple means of reducing cell or enzyme loss, and can improve the heat stability of entrapped enzymes. As alginates (particularly their calcium salts) form ionotropic gels spontaneously under low-temperature conditions, they are ideal for entrapment of whole cells or enzymes, which would otherwise be damaged under more stringent conditions (Navratil et al. 2002). However, as alginates or other viscous entrapment media act as a diffusion barrier, they may also reduce cell/enzyme access to the substrates or decrease the rates of release of end products. Therefore, this may result in a reduction in the efficiency of this type of process (Mishra and Kar 2003). This must be taken into account when choosing an immobilisation medium for large-scale industrial processes.
216
I.A. Brownlee et al.
While enzyme immobilisation is suitable for a number of applications, the use of whole cell immobilisation is necessary to carry out multistep, multienzyme reactions, especially those dependent on the presence of enzyme cofactors. Alginate is the most frequently described immobilisation agent in previous literature (Boyaval and Goulet 1988; Brandenberger and Widmer 1998). Table 2 shows some examples of alginate immobilisation technologies used in food production. One of the more recent developments in alginate use by the food industry has been in the encapsulation of live cells (probiotics) for delivery in foods to the human large bowel. A probiotic has been defined as: “A preparation of or a product containing viable, defined micro-organisms in sufficient numbers, which alter the microflora (by implantation or colonization) in a compartment of the host, and by that exert beneficial health effects in the host” (Schrezenmeir and De Vrese 2001). Probiotic cultures routinely contain lactic acid producing species such as lactobacillus, or bifidobacteria. While a number of benefits have been hypothesised for probiotic cultures in animal experiments and in vitro, there is still no clear-cut evidence of a beneficial effect on health of such preparations from human intervention studies (De Roos and Katan 2000; Allen et al. 2004; Honeycutt et al. 2007; Mallon et al. 2007). While the benefits to health of probiotics are likely to be a bone of contention for a considerable time, it is apparent that formulation of probiotic bacteria in foods
Table 2 Examples of whole-cell and enzyme alginate immobilisation technologies in the food industry Whole-cell application
Cells immobilised
Starter cultures for dairy industry Milk lactose hydrolysis (production of low-lactose milk) Production of lactic acid from cheese whey Production of fermented beverages Production of isomaltulose from sucrose
Various lactic acid producing Champagne et al. (1994) bacteria Kluyveromyces lactis Genari et al. (2003)
Bioindicators of thermal sterilisation Enzyme application Reduction of fruit juice bitterness Glucose determination of liquid foodstuffs Production of inverted sugar syrups Dextran synthesis Increasing n-3 polyunsaturated fatty acid concentrations in sardine Oil
Reference
Lactobacillus helveticus
Boyaval and Goulet (1988)
Various yeast cultures Protaminobacter rubrum Erwinia rhapontica Erwinia sp. D12 Serratia plymuthica Klebsiella planticola Klebsiella sp. Bacillus subtitles Bacillus stearothermophilus Enzymes immobilised
Navratil et al. (2002) Kawaguti et al. (2006)
Naringinase Glucose oxidase
Puri et al. (1996) Serp et al. (2002)
Inulinase Dextransucrases Lipase
Catana et al. (2005) De Segura et al. (2004) Okada and Morrissey (2008)
Serp et al. (2002) Reference
Applications of Alginates in Food
217
must ensure adequate delivery of viable cells to the large bowel for maximal effect. Simple and inexpensive technologies to form gel microcapsules from alginates under conditions which will not cause damage to the bacterial cells contained within are readily available. Alginates with a high G block content are the best alginates for microcapsule formation, owing to their high mechanical stability, high porosity and tolerance to salts and chelating agents (Kailasapathy 2002). Alginates act to increase survival of bacterial cells during food storage (Kailasapathy 2006; Krasaekoopt et al. 2006), and also reduce cell destruction under the inhospitable conditions of the stomach and small intestine (Kim et al. 2008; Ding and Shah 2007; Muthukumarasamy et al. 2006). In addition, encapsulation was shown to increase cell survival through a number of food processes, such as freezing (Kailasapathy 2002) and freeze-drying (Kailasapathy 2002; Capela et al. 2006). Microencapsulation has a number of further applications in the food industry, including encapsulation of reactive or volatile molecules, such as acidulants, fats and flavours (Desai and Park 2005). However, alginates are hydrophilic by nature, and therefore tend to be immiscible with these types of hydrophobic compound. Further from this, microencapsulation of low molecular weight components requires low matrix porosity so that entrapped compounds do not leach out. As alginates form highly porous gels, they may not be suitable for this type of application.
5
Food Coating
Edible films and coatings for foodstuffs are developing technologies that have high potential, owing to the current call for reduction or replacement of non-biodegradable or non-recyclable food packaging (Wang et al. 2007). As with conventional food packaging, these new technologies must heighten product stability, safety and shelf life. Alginates are one of a wide range of polysaccharides and proteins that have been used as an edible coating for a number of foodstuffs. Food coatings formed from sodium alginate have been shown to have excellent tensile strength, flexibility and resistance to tearing, and are impermeable to oils. However, owing to the porous nature of alginate gels, these coatings tend to have high permeability to oxygen and water (Wang et al. 2007). Alginate food coatings can also be formed ionotropically at room temperature, and therefore their use may be advantageous in a number of food applications. Incorporation of antimicrobial agents in the alginate gel has been demonstrated to be an effective barrier to microbial surface spoilage of vegetables, meat (Oussalah et al. 2007, 2006; Mehyar et al. 2007; Natrajan and Sheldon 2000) and fish products (Datta et al. 2008). The cold formation of the alginate gel coating reduces damage both to the antimicrobial entities and to the foodstuff itself. This property has also been demonstrated to be useful in the coating of a number of fresh fruit and vegetable products, such as lettuce (Tay and Perera 2004) and freshly cut apple and melon sections
218
I.A. Brownlee et al.
(Oms-Oliu et al. 2008; Rojas-Grau et al. 2008), where the coating increased shelf life, reduced browning, maintained crispness and texture and reduced vitamin C loss.
6
Physiological Effects of Dietary Alginates
As alginates are polysaccharides that are indigestible in the human stomach and small intestine, they would fit any of the variety of definitions that are available to describe dietary fibre (Brownlee et al. 2006; Champ et al. 2003; DeVries et al. 2001; McCleary 2003; Ha et al. 2000; Prosky 2000). As such, they exhibit many of the physiological characteristics commonly associated with dietary fibre, such as laxation and stool bulking. At the same time, alginates also exhibit a range of unique physiological actions. As with many of their food applications, the physiological effects of alginates are dependent on their biochemical and biophysical properties. Further understanding of health benefits of alginate ingestion, and particularly how these are associated with the biochemical and biophysical properties of alginates, is likely to open up new product opportunities for the food industry.
6.1
Alginate Safety
Alginates have been classified as safe for additive-level use both in the EU and in the USA. Previous studies into the dietary effects of sodium alginate and PGA have shown no adverse effects (i.e. no effect on haematological indices, plasma and urine biochemistry and breath hydrogen concentration) of ingestion of 175 mg kg−1 of body weight for 7 days followed by 200 mg kg−1 for a further 16 days in small groups of male volunteers (n = 5 for both compounds) (Anderson et al. 1991a, b).
6.2
Effect on Colonic Microflora
Incubation of human faecal inocula with alginates did not produce short-chain fatty acids for over 6 h (Michel et al. 1996). After 24-h incubation, between 50 and 80% of the alginate had been degraded under these conditions (Michel et al. 1996; Bobin-Dubigeon et al. 1997). These data suggest that alginates are slowly fermented by the colonic microflora in humans, although it must be noted that faecal inocula will not necessarily be indicative of the colonic microfloral population. In animal studies, the fermentability of dietary alginates increased over time of feeding, suggesting a shift of the colonic microflora to one that was more capable of degrading alginate polysaccharide chains (Suzuki et al. 1993a). From these previous studies, there is the suggestion that M-rich alginates are less well digested than G-rich ones, and that increasing chain length of alginates results in reduced fermentability (Michel et al. 1996, 1999; Suzuki et al. 1993a).
Applications of Alginates in Food
219
Owing to the complexity of the human colonic microflora, it is always difficult to predict how changes to specific bacterial species will affect the microflora as a whole, or the knock-on physiological effects these changes may have on the host. To date, very few data are available on the effects of alginates on bacterial species. The ability of 21 strains of authentic human intestinal bacteria to grow on alginatebased media has been tested under in vitro conditions. Only Bacteroides ovatus demonstrated the capability to thrive under these conditions (Fujii et al. 1992). One previous study where human participants were fed 10 g of alginates per day demonstrated an increase in faecal bifidobacterial cultures, and a decrease in both some potentially pathogenic bacterial strains (e.g. Enterobacteriaceae and lecithinasenegative Clostridia) and the levels of faecal toxins produced by putrefaction (e.g. ammonia and sulphides) (Terada et al. 1995). Table 3 compares the effects of a range of polysaccharides commonly used as thickeners, stabilisers and emulsifiers in the food industry on the colonic microflora.
6.3
Reduction of Toxicity of Colonic Luminal Contents
Bulking of the colonic contents (and eventually the passed stools) results in a dilution of any damaging agents in the colon, thereby effecting a reduced mucosal exposure to these factors. Seven-day supplementation of the diet of 5 healthy adult males with sodium alginate (175mg/kg body weight) significantly increased faecal wet and dry weights
Table 3 Effect of a range of viscous polysaccharides commonly used in the food industry on the colonic microflora. (Data collated from Bobin-Dubigeon et al. 1997; Fujii et al. 1992; Terada et al. 1995; Cherbut et al. 2003; Bliss et al. 1996; Barry et al. 1995; Bourquin et al. 1996; Ferguson and Jones 2000; Gulfi et al. 2005, 2006, 2007; Gibson et al. 1991; Macfarlane et al. 1998) Approximate percentage small-chain fatty acid production Fermentability (%) Acetate
Propionate
Alginate
50–80
66
19
Pectin
>83
77
8
Carrageenan 24
75
10
Guar gum Gum arabic
53 69
33 18
70–95 77
Small-chain fatty acid data typically do not add minor fatty acids in fermentation
Butyrate
Effects on microfloral strains
↑ Bacteroides ↑ Bifidobacteria ↓ Enterobacteriaceae ↓ Clostrida 14 ↑ Enterobacteriaceae ↑ Bacteroides 14 ↑ Sulphate-producing bacteria ↓ Bacteroides 14 ↑ Bifidobacteria 11 ↑ Bifidobacteria ↑ Lactobacilli up to 100% owing to the production of other, 15
220
I.A. Brownlee et al.
compared to outputs during a normal diet. A similar supplementation regime with PGA had no effect on these parameters (Anderson et al. 1991a, b). In porcine studies, 5% alginate in the feed increased the volume of colonic luminal contents. This effect was reported to be higher than that with cellulose, xylan and carageenan at the same concentration (Hoebler et al. 2000). A range of damaging agents which originate from food, microflora and the gastrointestinal tract itself that occur in the colon have been shown to be directly adsorbed to dietary alginates in a way similar to that of other dietary fibres (Brownlee et al. 2005).
6.4
Effect of Alginate on Satiety
Unlike other viscous polysaccharides, which need to be administered in gel form to be a gel in the stomach, alginates have the unique ability to form gels spontaneously at low temperature in the presence of acid or calcium ions. High viscosity or gel strength in the mouth is well associated with poor organoleptic acceptability of foods (Wolf et al. 2002). At the same time, high viscosity in the stomach is linked to increased gastric distension and thereby increased satiety (Hoad et al. 2004). Previous research has suggested that, as alginates can be administered in a low-viscosity form and gel spontaneously in the stomach, they represent a good candidate for formulations targeted at increasing satiety. In a participants study (n = 12) addition of either alginates or guar gum to a milk based liquid meal replacement resulted in participants reporting an increased feelings of fullness compared to a control (meal replacement only). Alginates with high gel strength significantly prolonged a postprandial feeling of hunger compared with both control and guar gum meal replacements. Alginate increased the volume of stomach contents, but did not affect gastric emptying rates (Hoad et al. 2004). A two-part beverage that formed a calcium alginate/pectin gel in the stomach was used in a participant study measuring dietary restraint in overweight and obese, non-dieting female volunteers (the beverage was taken twice daily). This beverage increased dietary restraint in participants with low initial dietary restraint, but not in those who already had high restraint levels. The beverage significantly reduced total energy intake, seemingly as a result of reducing carbohydrate intake (no significant effect was seen on total protein or fat intakes) (Pelkman et al. 2007). A snack bar formulation including alginate and guar gum had no reported effect on appetite compared with control snack bars (Mattes 2007).
6.5
Effect on Intestinal Absorption Rates
Alginates appear to have some inhibitory effects on a range of digestive enzymes in vitro. This is likely due to their high viscosity, which in turn reduces the availability of substrate for enzyme action. These effects are reviewed in detail elsewhere (Brownlee et al. 2005). As with other viscous fibres, inclusion of high levels of alginate in the diet has been linked with a reduced bioavailability of certain beneficial dietary components,
Applications of Alginates in Food
221
including b-carotene (Riedl et al. 1999), and minerals, such as calcium (Bosscher et al. 2001), iron, chromium and cobalt (Harmuth-Hoene and Schlenz 1980). This reduced nutrient and mineral absorption might suggest that the inclusion of high levels of alginate in the diets of at-risk individuals (e.g. the elderly, pregnant women and infants) may outweigh any potential health benefits. However, presaturation of alginate formulations with calcium and other divalent cations could reduce the possibility of these negative side effects.
6.6
Effect on Plasma Cholesterol
Increased intake of viscous dietary fibres has been linked to a reduction in plasma cholesterol concentrations (Dikeman and Fahey Jr. 2006). A meta-analysis of 67 controlled trials suggested that pectin had the greatest effect on lowering total cholesterol concentrations per gram of four common viscous fibre sources (reduction in cholesterol of 70 mM L−1 plasma g−1 fibre, compared with 37 mM L−1 g−1 for oat products, 28 mM L−1 g−1 for psyllium and 26 mM L−1 g−1 for guar gum). These effects appeared to be almost entirely due to a reduction in LDL cholesterol, rather than HDL cholesterol (Brown et al. 1999). This style of analysis does not account for variations in the study populations used, and currently no single study has adequately compared the cholesterol-lowering effects of dietary fibre types with each other within the same population. To date, relatively few studies have considered the effects of alginates in lowering cholesterol. Alginate (7.5 g day−1, M to G ratio 1.5) supplementation of a low-fibre diet has previously been shown to more than double (140% increase) mean fatty acid excretion in the digesta of a small cohort (n = 6) of human ileostomy patients (Sandberg et al. 1994). A number of animal model studies have also demonstrated that the presence of alginate in the small intestinal lumen decreases uptake of fats and reduces plasma cholesterol under a range of different diets (Brownlee et al. 2005; Jimenez-Escrig and Sanchez-Muniz 2000). These effects are likely to be due to the increased levels of faecal bile and cholesterol excretion that have previously been reported (Seal and Mathers 2001; Kimura et al. 1996). Levels of 1 and 3% sodium alginate were shown to have hypocholesteraemic effects (mean reduction in plasma cholesterol from fibre-free controls of 8.5 and 20.5%, respectively) similar to those of the algal polysaccharide funoran (7.3 and 20.9%), but less than those of carrageenan (14.6 and 29.9%) (Jimenez-Escrig and Sanchez-Muniz 2000; Ito and Tsuchiya 1972) in rats fed 1% dietary cholesterol. All of these fibre types had a much larger effect than agar (3% in diet gave a reduction of only 1.8% in total plasma cholesterol). Within this study, it was also shown that low molecular weight alginates do not appear to have these hypocholesteraemic effects. In rat diets with higher total cholesterol and fat content, sodium alginate inclusion had effects on total cholesterol reduction similar to those of inclusion of a range of other algal polysaccharides (sulphated glucuronoxylohamnan, porphyran and furonan), and reduced blood cholesterol concentrations more than fucoidan and agar (Jimenez-
222
I.A. Brownlee et al.
Escrig and Sanchez-Muniz 2000; Ren et al. 1994). Feeding of a cholesterol-free diet to rats containing 5 or 10% sodium alginate or guar gum caused a similar drop in plasma cholesterol concentrations over a 21-day period (Seal and Mathers 2001). Increased G-block content will lead to an increased gel-forming capacity of the alginate, which would be expected to reduce intestinal absorption. The effects of alginate on cholesterol uptake into the bloodstream have previously been suggested to be increased with a lower M to G ratio (Suzuki et al. 1993b), but this was believed to be due to the effects of the high-viscosity alginate on depression of appetite (i.e. lower levels of ingested cholesterol), rather than effects on intestinal absorption and/or increased cholesterol excretion.
6.7
Effect on Glycaemic and Insulinaemic Responses
Glucose absorption rates have also been shown to be reduced in the presence of alginate. A 5-g supplement of sodium alginate to test meals containing similar levels of digestible carbohydrates, fats and proteins was given to a cohort of diabetes type II patients. This low level of dietary alginate caused a reduction in blood peak glucose and plasma insulin rise (by 31 and 42%, respectively), as well as prolongation of postprandial emptying (Torsdottir et al. 1991). Similar results of blunting postprandial plasma glucose and insulin elevation have also been reported upon feeding a low dose of alginate (1.5 g sodium alginate within a liquid drink) to healthy human subjects (Wolf et al. 2002). Within this study, the 1.5 g alginate drink reduced both peak postprandial glucose concentrations and total glucose uptake over 3 h significantly more than a comparable drink containing 1.2 g of a mixture of gum arabic and 0.3 g of guar gum in the same population of 30 healthy adults. Snack bars including alginate also reduced postprandial peak glucose concentrations and total glucose uptake (over 3 h) when compared with snack bars containing guar gum (Williams et al. 2004).
7
Summary
Alginates are being used or have the potential to be used for an increasing variety of applications in the food industry. Their unique gelling abilities at low temperature alongside good heat stability make them ideal for use as thickeners, stabilisers or restructuring agents. In addition to this, alginates also have uses in a myriad of newer food applications, from encapsulating active enzymes and live bacteria, to acting as the carrier for protective coating of prepacked, cut or prepared fruits and vegetables. A wide range of alginates are commercially available with various bioactive properties, mainly dependent on polysaccharide chain length and the ratio of mannuronate to guluronate sugars, thus allowing for a large scope of potential uses. Alginate intake in the diet has a number of physiological effects both in the gastrointestinal tract and systemically that are relatively similar to those of a range of
Applications of Alginates in Food
223
other viscous polysaccharides. Once again, the unique biophysical properties of alginates separate them from other similar dietary fibres. Alginates can be given in lowviscosity formulations, which firstly allows both increased palatability and the possibility of a higher oral dose of alginates being delivered than is possible for other viscous polysaccharides. Secondly, as alginate formulations gel in the stomach, in the presence of either acid or calcium ions, they can be used to increase satiety and reduce absorption and digestion rates of other macronutrients in the diet. Although alginates have been demonstrated to be beneficial in this way in a number of human participant studies, current legislation worldwide only allows for their use in food as an additive (as opposed to an ingredient). This means that alginate used in foods will remain at a low concentration for the foreseeable future. However, this still allows the possibility of novel applications of specific alginates in the food industry that have high bioactivity at low concentration. Therefore, food manufacturers, alginate producers and food and nutrition scientists will need to work more closely together if novel uses for these unique algal and bacterial polysaccharides are to be identified.
References Allen SJ, Okoko B, Martinez E, Gregorio G, Dans LF (2004) Probiotics for treating infectious diarrhoea. Cochrane Database Syst Rev (2):CD003048 Anderson DMW, Brydon WG, Eastwood MA, Sedgwick DM (1991a) Dietary effects of propylene glycol alginate in humans. Food Addit Contam 8:225–236 Anderson DMW, Brydon WG, Eastwood MA, Sedgwick DM. (1991b) Dietary effects of sodium alginate in humans. Food Addit Contam 8:237–248 Baeza R, Sanchez CC, Pilosof AMR, Patino JMR (2004) Interfacial and foaming properties of prolylenglycol alginates: effect of degree of esterification and molecular weight. Colloids Surf B Biointerfaces 36:139–145 Barry JL, Hoebler C, Macfarlane GT, Macfarlane S, Mathers JC, Reed KA, Mortensen PB, Nordgard I, Rowland IR, Rumney CJ (1995) Estimation of the fermentability of dietary fibre in vitro: a European interlaboratory study. Br J Nutr 74:303–322 Bliss DZ, Stein TP, Schleifer CR, Settle RG (1996) Supplementation with gum arabic fiber increases fecal nitrogen excretion and lowers serum urea nitrogen concentration in chronic renal failure patients consuming a low-protein diet. Am J Clin Nutr 63:392–398 Bobin-Dubigeon C, Hoebler C, Lognone V, Dagorn-Scaviner C, Mabeau S, Barry JL, Lahaye M (1997) Chemical composition, physico-chemical properties, enzymatic inhibition and fermentative characteristics of dietary fibres from edible seaweeds. Sci Des Aliments 17:619–639 Boles JA, Shand PJ (1999) Effects of raw binder system, meat cut and prior freezing on restructured beef. Meat Sci 53:233–239 Bosscher D, Van Caillie-Bertrand M, Deelstra H (2001) Effect of thickening agents, based on soluble dietary fiber, on the availability of calcium, iron, and zinc from infant formulas. Nutrition 17:614–618 Bourquin LD, Titgemeyer EC, Fahey GC (1996) Fermentation of various dietary fiber sources by human fecal bacteria. Nutr Res 16:1191–1131 Boyaval P, Goulet J (1988) Optimal conditions for production of lactic acid from cheese whey permeate by Ca-alginate-entrapped Lactobacillus helveticus. Enzyme Microb Technol 10:725–728 Brandenberger H, Widmer F (1998) A new multinozzle encapsulation/immobilisation system to produce uniform beads of alginate. J Biotechnol 63:73–80
224
I.A. Brownlee et al.
Brown L, Rosner B, Willett WW, Sacks FM (1999) Cholesterol-lowering effects of dietary fiber: a meta-analysis. Am J Clin Nutr 69:30–42 Brownlee IA, Allen A, Pearson JP, Dettmar PW, Havler ME, Atherton MR, Onsoyen E. (2005) Alginate as a source of dietary fiber. Crit Rev Food Sci Nutr 45:497–510 Brownlee IA, Dettmar PW, Strugala V, Pearson JP (2006) The interaction of dietary fibres with the colon. Curr Nutr Food Sci 2:243–64 Bryan JL, Little SL, Sykes J, Baxter T, Dettmar PW (2001) Efficacy and safety of a unique antireflux agent, Gaviscon advance, for the treatment of heartburn in pregnancy. Gastroenterology 120:2214 Capela P, Hay TKC, Shah NP (2006) Effect of cryoprotectants, prebiotics and microencapsulation on survival of probiotic organisms in yoghurt and freeze-dried yoghurt. Food Res Int 39:203–211 Catana R, Ferreira BS, Cabral JMS, Fernandes P (2005) Immobilization of inulinase for sucrose hydrolysis. Food Chem 91:517–520 Champ M, Langkilde A-M, Brouns F, Kettlitz B, Collet YLB (2003) Advances in dietary fibre characterisation. 1. Definition of dietary fibre, physiological relevance, health benefits and analytical aspects. Nutr Res Rev 16:71–82 Champagne CP, Lacroix C, Sodini-Gallot I (1994) Immobilized cell technologies for the dairy industry. Crit Rev Biotechnol 14:109–134 Cherbut C, Michel C, Raison V, Kravtchenko TP, Severine M (2003) Acacia gum is a bifidogenic dietary fibre with high digestive tolerance in healthy humans. Microb Ecol Health Dis 15:43–50 Cong-Gui C, Gerelt B, Shao-Tong J, Nishiumi T, Suzuki A (2006) Effects of high pressure on pH, water-binding capacity and textural properties of pork muscle gels containing various levels of sodium alginate. Asian-Australas J Anim Sci 19:1658–1664 Das Gupta DK, Roopa N, Leela RK (2007) Development of stable restructured mango gel. Am J Food Technol 2:176–182 Datta S, Janes ME, Xue Q-G, Losso J, La Peyre JF (2008) Control of Listeria monocytogenes and Salmonella anatum on the surface of smoked salmon coated with calcium alginate coating containing oyster lysozyme and nisin. J Food Sci 73:M67–M71 De Roos NM, Katan MB (2000) Effects of probiotic bacteria on diarrhea, lipid metabolism, and carcinogenesis: a review of papers published between 1988 and 1998. Am J Clin Nutr 71:405–411 De Ruiter GA, Rudolph B (1997) Carrageenan biotechnology. Trends Food Sci Technol 8:389–395 Desai KGH, Park HJ (2005) Recent developments in microencapsulation of food ingredients. Drying Technol 23:1361–1394 De Segura AG, Alcalde M, Lopez-Cortes N, Plou FJ, Ballesteros A (2004) Modulating the synthesis of dextran with the acceptor reaction using native and encapsulated dextransucrases. Food Technol Biotechnol 42:337–342 de Vos P, Faas MM, Strand B, Calafiore R (2006) Alginate-based microcapsules for immunoisolation of pancreatic islets. Biomaterials 27:5603–5617 Devatkal S, Mendiratta SK (2001) Use of calcium lactate with salt-phosphate and alginate-calcium gels in restructured pork rolls. Meat Sci 58:371–379 de Vries J (2004) Hydrocolloid gelling agents and their applications. In: Williams PA, Phillips GO (eds) Gums and stabilisers for the food industry – 12. IRL, Oxford, pp 23–31 DeVries JW, Camire ME, Cho S, Craig S, Gordon D, Jones JM, Li B, Lineback D, Prosky L, Tungland BC (2001) The definition of dietary fiber. Cereal Foods World 46:112–129 Dikeman CL, Fahey GC Jr (2006) Viscosity as related to dietary fiber: a review. Crit Rev Food Sci Nutr 46:649–663 Ding WK, Shah NP (2007) Acid, bile, and heat tolerance of free and microencapsulated probiotic bacteria. J Food Sci 72:M446–M450 Ertesvag H, Valla S (1998) Biosynthesis and applications of alginates. Polym Degrad Stabil 59:85–91 Ferguson MJ, Jones GP (2000) Production of short-chain fatty acids following in vitro fermentation of saccharides, saccharide esters, fructo-oligosaccharides, starches, modified starches and non-starch polysaccharides. J Sci Food Agric 80:166–170
Applications of Alginates in Food
225
Ferreira IMPLV, Jorge K, Nogueira LC, Silva F, Trugo LC. (2005) Effects of the combination of hydrophobic polypeptides, iso-alpha acids, and malto-oligosaccharides on beer foam stability. J Agric Food Chem 53:4976–4981 Food Standards Agency (2002) Food additives legislation: guidance notes. Food Standards Agency, London Fujii T, Kuda T, Saheki K, Okuzumi M. (1992) Fermentation of water-soluble polysaccharides of brown-algae by human intestinal bacteria invitro. Nippon Suisan Gakk 58:147–152 Gaonkar AG (1991) Surface and interfacial activities and emulsion characteristics of some food hydrocolloids. Food Hydrocolloids 5:329–337 Genari AN, Passos FV, Passos FML (2003) Configuration of a bioreactor for milk lactose hydrolysis. J Dairy Sci 86:2783–2789 George M, Abraham TE (2006) Polyionic hydrocolloids for the intestinal delivery of protein drugs: Alginate and chitosan – a review. J Control Release 114:1–14 Gibson GR, Cummings JH, Macfarlane GT (1991) Growth and activities of sulphate-reducing bacteria in gut contents of healthy subjects and patients with ulcerative colitis. FEMS Microbiol Ecol 86:103–111 Gomez-Diaz D, Navaza JM (2004) Rheology of food stabilizers blends. J Food Eng 64:143–149 Grizotto RK, Berbari SAG, De Moura SCSR, Claus ML (2006) Estudo da vida-de-prateleira de fruta estruturada e desidratada obtida de polpa concentrada de mamao (Shelf life studies of restructured and dried fruit made from concentrated papaya pulp). Cienc Tecnol Aliment 26:709–714 Grizotto RK, Bruns RE, De Aguirre JM, De Menezes HC (2007) Technological aspects for restructuring concentrated pineapple pulp. LWT Food Sci Technol 40:759–765 Groboillot A, Boadi DK, Poncelet D, Neufeld RJ (1994) Immobilization of cells for application in the food industry. Crit Rev Biotechnol 14:75–107 Gujral HS, Sharma P, Singh N, Sogi DS (2001) Effect of hydrocolloids on the rheology of tamarind sauce. J Food Sci Technol 38:314–318 Gulfi M, Arrigoni E, Amado R (2005) Influence of structure on in vitro fermentability of commercial pectins and partially hydrolysed pectin preparations. Carbohydr Polym 59:247–255 Gulfi M, Arrigoni E, Amado R (2006) The chemical characteristics of apple pectin influence its fermentability in vitro. LWT Food Sci Technol 39:1001–1004 Gulfi M, Arrigoni E, Amado R (2007) In vitro fermentability of a pectin fraction rich in hairy regions. Carbohydr Polym 67:410–416 Ha M-A, Jarvis MC, Mann JI (2000) A definition for dietary fibre. Eur J Clin Nutr 54:861–864 Harmuth-Hoene A-E, Schlenz R (1980) Effect of dietary fiber on mineral absorption in growing rats. J Nutr 110:1774–1784 Hoad CL, Rayment P, Spiller RC, Marciani L, De Celis Alonso B, Traynor C, Mela DJ, Peters HPF, Gowland PA (2004) In vivo imaging of intragastric gelation and its effect on satiety in humans. J Nutr 134:2293–2300 Hoebler C, Guillon F, Darcy-Vrillon B, Vaugelade P, Lahaye M, Worthington E, Duee PH, Barry JL (2000) Supplementation of pig diet with algal fibre changes the chemical and physicochemical characteristics of digesta. J Sci Food Agric 80:1357–1364 Honeycutt TCB, El Khashab M, Wardrop III RM, McNeal-Trice K, Honeycutt ALB, Christy CG, Mistry K, Harris BD, Meliones JN, Kocis KC (2007) Probiotic administration and the incidence of nosocomial infection in pediatric intensive care: A randomized placebo-controlled trial. Pediatr Crit Care Med 8:452–458 Huang X, Kakuda Y, Cui W (2001) Hydrocolloids in emulsions: particle size distribution and interfacial activity. Food Hydrocolloids 15:533–542 Hubbermann EM, Heins A, Stockmann H, Schwarz K (2006) Influence of acids, salt, sugars and hydrocolloids on the colour stability of anthocyanin rich black currant and elderberry concentrates. Eur Food Res Technol 223:83–90 Ingram N, Wright TA, Ingoldby CJH (1998) A prospective randomized study of calcium alginate (Sorbsan) versus standard gauze packing following haemorrhoidectomy. J R Coll Surg Edinb 43:308–309
226
I.A. Brownlee et al.
Ito K, Tsuchiya Y (1972) The effect of algal polysaccharides on the depressing of plasma cholesterol levels in rats. Paper presented at the seventh international seaweed symposium, Sappora, Japan Jensen A (1993) Present and future needs for algae and algal polysaccharides. Hydrobiologia 260/261:15–23 Jimenez-Escrig A, Sanchez-Muniz FJ (2000) Dietary fibre from edible seaweeds: chemical structure, physicochemical properties and effects on cholesterol metabolism. Nutr Res 20:585–598 Kailasapathy K (2002) Microencapsulation of probiotic bacteria: technology and potential applications. Curr Issues Intest Microbiol 3:39–48 Kailasapathy K (2006) Survival of free and encapsulated probiotic bacteria and their effect on the sensory properties of yoghurt. LWT Food Sci Technol 39:1221–1227 Kawaguti HY, Manrich E, Sato HH (2006) Production of isomaltulose using Erwinia sp. D12 cells: culture medium optimization and cell immobilization in alginate. Biochem Eng J 29:270–277 Kim S-J, Cho SY, Kim SH, Song OJ, Shin IS, Cha DS, Park HJ (2008) Effect of microencapsulation on viability and other characteristics in Lactobacillus acidophilus ATCC 43121. LWT Food Sci Technol 41:493–500 Kimura Y, Watanabe K, Okuda H (1996) Effects of soluble sodium alginate on cholesterol excretion and glucose tolerance in rats. J Ethnopharmacol 54:47–54 Krasaekoopt W, Bhandari B, Deeth HC (2006) Survival of probiotics encapsulated in chitosancoated alginate beads in yoghurt from UHT- and conventionally treated milk during storage. LWT Food Sci Technol 39:177–183 Kumar M, Sharma BD, Kumar RR (2007) Evaluation of sodium alignate as a fat replacer on processing and shelf-life of low-fat ground pork patties. Asian-Australas J Anim Sci 20:588–597 Kumar N, Sahoo J (2006) Studies on use of sodium alginate as fat replacer in development of low-fat chevon loaves. J Food Sci Technol 43:410–412 Lee DW, Hwang SJ, Park JB, Park HJ (2003) Preparation and release characteristics of polymercoated and blended alginate microspheres. J Microencapsul 20:179–192 Lin KW, Keeton JT (1998) Textural and physicochemical properties of low-fat, precooked ground beef patties containing carrageenan and sodium alginate. J Food Sci 63:571–574 Macfarlane S, Quigley ME, Hopkins MJ, Newton DF, Macfarlane GT (1998) Polysaccharide degradation by human intestinal bacteria during growth under multi-substrate limiting conditions in a three-stage continuous culture system. FEMS Microbiol Ecol 26:231–243 Mallon P, McKay D, Kirk S, Gardiner K (2007) Probiotics for induction of remission in ulcerative colitis. Cochrane Database Syst Rev (4):CD005573 Mancini F, McHugh TH (2000) Fruit-alginate interactions in novel restructured products. Nahrung 44:152–157 Mancini F, Montanari L, Peressini D, Fantozzi P (2002) Influence of alginate concentration and molecular weight on functional properties of mayonnaise. LWT Food Sci Technol 35:517–525 Mandel KG, Daggy BP, Brodie DA, Jacoby HI (2000) Review article: alginate-raft formulations in the treatment of heartburn and acid reflux. Aliment Pharmacol Ther 14:669–690 Manjunatha SS, Gupta DKD (2006) Instrumental textural characteristics of restructured carrot cubes. Int J Food Properties 9:453–462 Mattes RD (2007) Effects of a combination fiber system on appetite and energy intake in overweight humans. Physiol Behav 90:705–711 McCleary BV (2003) Dietary fibre analysis. Proc Nutr Soc 62:3–9 Mehyar GF, Han JH, Holley RA, Blank G, Hydamaka A. (2007) Suitability of pea starch and calcium alginate as antimicrobial coatings on chicken skin. Poult Sci 86:386–393 Michel C, Lahaye M, Bonnet C, Mabeau S, Barry JL (1996) In vitro fermentation by human faecal bacteria of total and purified dietary fibres from brown seaweeds. Br J Nutr 75:263–280 Michel C, Benard C, Lahaye M, Formaglio D, Kaeffer B, Quemener B, Berot S, Yvin JC, Blottiere HM, Cherbut C (1999) Algal oligosaccharides as functional foods: in vitro study of their cellular and fermentative effects. Sci Des Aliments 19:311–332 Mishra P, Kar R (2003) Treatment of grapefruit juice for bitterness removal by amberlite IR 120 and amberlite IR 400 and alginate entrapped naringinase enzyme. J Food Sci 68:1229–1233
Applications of Alginates in Food
227
Moe S, Draget KI, Skjak-Braek G, Smidsrod O (1995) Alginates. In: Stephen AM (ed) Food polysaccharides and their applications. Decker, New York, pp 245–86 Muthukumarasamy P, Allan-Wojtas P, Holley RA (2006) Stability of Lactobacillus reuteri in different types of microcapsules. J Food Sci 71:M20–M24 Natrajan N, Sheldon BW (2000) Inhibition of Salmonella on poultry skin using protein- and polysaccharide-based films containing a nisin formulation. J Food Prot 63:1268–1272 Navratil M, Gemeiner P, Klein J, Sturdik E, Malovikova A, Nahalka J, Vikartovska A, Domeny Z, Smogrovicova D (2002) Properties of hydrogel materials used for entrapment of microbial cells in production of fermented beverages. Artif Cells Blood Substit Immobil Biotechnol 30:199–218 Norton S, Vuillemard J-C (1994) Food bioconversions and metabolite production using immobilized cell technology. Crit Rev Biotechnol 14:193–224 Okada T, Morrissey MT (2008) Production of n-3 polyunsaturated fatty acid concentrate from sardine oil by immobilized Candida rugosa lipase. J Food Sci 73:C146–C150 Oms-Oliu G, Soliva-Fortuny R, Martin-Belloso O (2008) Using polysaccharide-based edible coatings to enhance quality and antioxidant properties of fresh-cut melon. LWT Food Sci Technol 41:1862–1870 Oussalah M, Caillet S, Salmieri S, Saucier L, Lacroix M (2006) Antimicrobial effects of alginatebased film containing essential oils for the preservation of whole beef muscle. J Food Prot 69:2364–2369 Oussalah M, Caillet S, Salmieri S, Saucier L, Lacroix M (2007) Antimicrobial effects of alginatebased films containing essential oils on Listeria monocytogenes and Salmonella typhimurium present in bologna and ham. J Food Prot 70:901–908 Paraskevopoulou A, Boskou D, Kiosseoglou V (2005) Stabilization of olive oil–lemon juice emulsion with polysaccharides. Food Chem 90:627–634 Paraskevopoulou D, Boskou D, Paraskevopoulou A (2006) Oxidative stability of olive oil–lemon juice salad dressings stabilized with polysaccharides. Food Chem 101:1197–1204 Pelkman CL, Navia JL, Miller AE, Pohle RJ (2007) Novel calcium-gelled, alginate-pectin beverage reduced energy intake in nondieting overweight and obese women: Interactions with dietary restraint status. Am J Clin Nutr 86:1595–1602 Prosky L (2000) What is dietary fiber? J AOAC Int 83:985–987 Puri M, Marwaha SS, Kothari RM (1996) Studies on the applicability of alginate-entrapped naringinase for the debittering of kinnow juice. Enzyme Microb Technol 18:281–285 Ranken MD (2000) Handbook of meat product technology. Blackwell, Oxford Regand A, Goff HD (2003) Structure and ice recrystallization in frozen stabilized ice cream model systems. Food Hydrocolloids 17:95–102 Ren D, Noda H, Amano H, Nishino T, Nishizana K (1994) Study on the hypertensive and antihyperlipidemic effect of marine algae. Fish Sci 60:83–88 Riedl J, Linseisen J, Hoffmann J, Wolfram G (1999) Some dietary fibers reduce the absorption of carotenoids in women. J Nutr 129:2170–2176 Rojas-Grau MA, Tapia MS, Martin-Belloso O (2008) Using polysaccharide-based edible coatings to maintain quality of fresh-cut Fuji apples. LWT Food Sci Technol 41:139–147 Sandberg AS, Andersson H, Bosaeus I, Carlsson NG, Hasselblad K, Harrod M (1994) Alginate, small-bowel sterol excretion, and absorption of nutrients in ileostomy subjects. Am J Clin Nutr 60:751–756 Sarker DK, Wilde PJ (1999) Restoration of protein foam stability through electrostatic propylene glycol alginate-mediated protein-protein interactions. Colloids Surf B Biointerfaces 15:203–213 Schrezenmeir J, De Vrese M (2001) Probiotics, prebiotics, and synbiotics – approaching a definition. Am J Clin Nutr 73:361S–364S Seal CJ, Mathers JC (2001) Comparative gastrointestinal and plasma cholesterol responses of rats fed on cholesterol-free diets supplemented with guar gum and sodium alginate. Br J Nutr 85:317–324 Serp D, Von Stockar U, Marison IW (2002) Immobilized bacterial spores for use as bioindicators in the validation of thermal sterilization processes. J Food Prot 65:1134–1141
228
I.A. Brownlee et al.
Skaugrud O, Hagen A, Borgersen B, Dornish M (1999) Biomedical and pharmaceutical applications of alginate and chitosan. Biotechnol Genet Eng Rev 16:23–40 Smidsrod O, Draget KI (1996) Chemistry and physical properties of alginates. Carbohydr Eur 14:7–13 Smith J, LilyHong S (2003) Food additives data book. Blackwell, Oxford Stokke BT, Draget KI, Smidsrod O, Yuguchi Y, Urakawa H, Kajiwara K (2000) Small-angle X-ray scattering and rheological characterization of alginate gels 1. Ca-alginate gels. Macromolecules 33:1853–1863 Suklim K, Flick GJ Jr, Marcy JE, Eigel WN, Haugh CG, Granata LA (2004) Effect of cold-set binders: Alginates and microbial transglutaminase on the physical properties of restructured scallops. J Texture Stud 35:634–42 Suzuki T, Nakai K, Yoshie Y, Shirai T, Hirano T (1993a) Digestibility of dietary fiber in brown alga, Kombu, by rats. Nippon Suisan Gakkaishi 59:879–84 Suzuki T, Nakai K, Yoshie Y, Shirai T, Hirano T (1993b) Effects of sodium alginates rich in guluronic and mannuronic acids on cholesterol levels and digestive organs of high-cholesterolfed rats. Nippon Suisan Gakkaishi 59:545 Sworn G (2004) Hydrocolloid thickeners and their applications. In: Williams PA, Phillips GO (eds) Gums and stabilisers for the food industry – 12. IRL, Oxford, pp 13–22 Tay SL, Perera CO (2004) Effect of 1-methylcyclopropene treatment and edible coatings on the quality of minimally processed lettuce. J Food Sci 69:FCT131–FCT135 Terada A, Hara H, Mitsuoka T (1995) Effect of dietary alginate on the faecal microbiota and faecal metabolic activity in humans. Microb Ecol Health Dis 8:259–266 Torsdottir I, Alpsten M, Holm G, Sandberg AS, Tolli J (1991) A small dose of soluble alginatefiber affects postprandial glycemia and gastric emptying in humans with diabetes. J Nutr 121:795–799 Uludag H, De Vos P, Tresco PA (2000) Technology of mammalian cell encapsulation. Adv Drug Deliv Rev 42:29–64 Wang LZ, Liu L, Holmes J, Kerry JF, Kerry JP (2007) Assessment of film-forming potential and properties of protein and polysaccharide-based biopolymer films. Int J Food Sci Technol 42: 1128–1138 Ward FM (2004) Water-soluble gums in low mositure foods (extruded snacks, noodles, and bakery products). In: Williams PA, Phillips GO (eds) Gums and stabilisers for the food industry – 12. IRL, Oxford, pp 484–495 Wendin K, Aaby K, Edris A, Ellekjaer MR, Albin R, Bergenstahl B, Johansson L, Willers EP, Solheim R (1997) Low-fat mayonnaise: Influences of fat content, aroma compounds and thickeners. Food Hydrocolloids 11:87–99 Williams JA, Lai C-S, Corwin H, Ma Y, Maki KC, Garleb KA, Wolf BW. (2004) Inclusion of guar gum and alginate into a crispy bar improves postprandial glycemia in human. J Nutr 134:886–889 Wolf BW, Lai CS, Kipnes MS, Ataya DG, Wheeler KB, Zinker BA, Garleb KA, Firkins JL. (2002) Glycemic and insulinemic responses of nondiabetic healthy adult subjects to an experimental acid-induced viscosity complex incorporated into a glucose beverage. Nutrition 18:621–626 Yamamoto K, Kumagai H, Sakiyama T, Song CM, Yano T (1992) Inhibitory activity of alginates against the formation of calcium-phosphate. Biosci Biotechnol Biochem 56:90–93
Alginate and Its Comonomer Mannuronic Acid: Medical Relevance as Drugs Abbas Mirshafiey and Bernd H.A. Rehm
Contents 1 Introduction ...................................................................................................................... 2 Therapeutic Effects of Alginate on Experimental Acute Ulcerative Colitis .................... 3 Treatment of Experimental Chronic Ulcerative Colitis with Sodium Alginate ............... 4 Treatment of Experimental Immune Complex Glomerulonephritis by Sodium Alginate ......................................................................................................... 5 Therapeutic Effects of M2000 on an Experimental Model of Rheumatoid Arthritis ...... 6 Immunosuppressive Effect of M2000 in Experimental Multiple Sclerosis ..................... 7 Treatment of Experimental Nephrotic Syndrome with M2000 ....................................... 8 Therapeutic Effects of M2000 in Experimental Immune Complex Glomerulonephritis........................................................................................... 9 Tolerability and Pharmacotoxicology of M2000 ............................................................. 10 Outlook ............................................................................................................................ References ................................................................................................................................
230 231 236 237 242 244 248 252 255 258 258
Abstract The therapeutic and anti-inflammatory properties of sodium alginate and one of its components, b-d-mannuronic acid, were tested in various experimental models. On the basis of its chemical structure, b-d-mannuronic acid had been conceived as an anti-inflammatory drug. b-d-Mannuronic acid is a constituent of alginate but is not available commercially; hence, polymannuronate was obtained from an epimerase (AlgG) negative mutant of Pseudomonas sp. and the monomer was isolated by acid hydrolysis. The oral and intraperitoneal administration of alginate gels in experimental models of ulcerative colitis and glomerulonephritis as well as the oral administration and intraperitoneal injections of b-d-mannuronic acid in different models of rheumatoid arthritis, multiple sclerosis, glomerulonephritis, and nephrotic syndrome showed that alginate as well as b-d-mannuronic acid are able to exhibit A. Mirshafiey Department of Immunology, School of Public Health, Tehran University of Medical Sciences, Tehran-14155, Box: 6446, Iran e-mail: [email protected] B.H.A. Rehm Institute of Molecular BioSciences, Massey University, Palmerston North, New Zealand, e-mail: [email protected]
B.H.A. Rehm (ed.), Alginates: Biology and Applications, Microbiology Monographs 13, DOI: 10.1007/978-3-540-92679-5_10, © Springer-Verlag Berlin Heidelberg 2009
229
230
A. Mirshafiey and B.H.A. Rehm
their therapeutic efficacy in inflammatory diseases. Recent experimental evidence has revealed that b-d-mannuronic acid represents a novel nonsteroidal anti-inflammatory drug with a low molecular weight and which provides a new therapeutic option in attenuation of inflammatory reactions and autoimmune diseases.
1
Introduction
Alginates are natural macromolecules composed of b- d -mannuronate and a- l-guluronate linked by 1→4 glycosidic bonds. They are synthesized by bacteria and/or brown seaweeds, and the mannuronate residues of the bacteria, but not seaweed polymers, are acetylated at positions 0–2 and/or 0–3 to a variable extent (Skjak-Braek et al. 1986; Bucke 1987; Rehm 1998). The variability in monomer block structures and acetylation which are associated with the source of alginate strongly affect the physicochemical and rheological properties of the alginate polymer (Rehm and Valla 1997). The alginate gels are well known as being biocompatible, degradable, and nontoxic; thus, they are widely used as carriers for drug delivery, hemostatic wound dressing, and immunoisolation systems for transplantation using uncoated alginate microspheres and devices anastomosed to the vascular system as arteriovenous shunts such as alginate-impregnated polyester vascular graft (Odell et al. 1994; Agren 1996; Murata et al. 2007; Lee et al. 1997; Efentakis and Buckton 2002; Li et al. 2006; Selmi et al. 2008; Sevgi et al. 2008). On the other hand, inhibitory effects of various types of alginic acid on hyaluronidase and mast cell degranulation were examined, and it was found that alginic acid with a mannuronate to guluronate ratio of 1.0 exhibited the strongest inhibition of both activities (Asada et al. 1997). Moreover, the protective and reparative effects of sodium alginate on radiation stomatitis and suppression of radioactive absorption by this compound in animals and human subjects were investigated (Hasegawa et al. 1989; Oshitani et al. 1990; Gong et al. 1991; McGlashan et al. 2009). The effect of alginic acid on an allergy model was analyzed by anaphylaxis, a histidine decarboxylase assay, and a histamine assay. Alginic acid dose dependently inhibited systemic anaphylaxis and passive cutaneous anaphylaxis and also decreased histamine release from serum and peritoneal mast cells. All these effects were stronger than those of disodium cromoglycate, the reference drug tested. In addition, alginic acid significantly inhibited the expression level of nuclear factor (NF)-kB and some proinflammatory cytokines, such as interleukin-1b and tumor necrosis factor a (TNF-a) (Jeong et al. 2006). Although (Yang and Jones 2008) have reported that sodium alginate causes innate immune responses through NF-kB activation, the majority of investigations have demonstrated the biocompatibility and anti-inflammatory effects of alginate and one of its components, b-d-mannuronic acid (M2000), in various experimental models and/or patients (Siebers et al. 1997; Kammerlander and Eberlein 2003; Ohsumi et al. 2005; Hori et al. 2008). In this chapter, our aim is to show the
Alginate and Its Comonomer Mannuronic Acid
231
tolerability and therapeutic potency of alginate and M2000 in different disease models as well as various pharmaceutically relevant investigations.
2
Therapeutic Effects of Alginate on Experimental Acute Ulcerative Colitis
The potential therapeutic effect of low-viscosity sodium alginate (LVA) was studied in a rat model of acute colitis induced by intracolonic administration of acetic acid (Mirshafiey et al. 2005a). To induce the experimental model of acute ulcerative colitis, the method of (Morris et al. 1989) and (Yamada et al. 1991) was used. This experimental model produced a significant ulcerative colitis. Induction of colitis also significantly enhanced the levels of serum and colonic mucosal cytokines (interleukin-6, IL-6; and TNF-a) and eicosanoids (leukotriene B4, LTB4; and prostaglandin E2, PGE2), which paralleled the severity of colitis. For treatment, LVA was dissolved in water and adjusted to a concentration of 0.5% (w/v). The prophylactic and therapeutic rat groups received 0.5% LVA solution as drinking water ad libitum for 7 days. The onset of oral LVA administration for prophylactic and therapeutic groups was 24 h before and after the induction of colitis, respectively. The rats were killed at the end of the week. The results showed that intracolonic administration of 4% acetic acid resulted in acute inflammatory bowel disease (IBD) in control nontreated rats. The control rats developed colonic macroscopic damage such as diffuse hyperemia and ulcerations. Daily oral treatment with LVA (drinking 35.5 ± 2.8 ml of 0.5% w/v LVA solution per rat, which was equivalent 0.65–0.77 mg kg−1 LVA per day) significantly reduced the colonic damage score in prophylactic and therapeutic rat groups (with mean scores of 5.0 and 4.6, respectively) when compared with control rats (mean score 9.4; P < 0.05). The microscopic examination of colonic specimens revealed a marked infiltration of inflammatory cells into both mucosa and submucosa, ulceration, and fibrosis. The most pronounced lesions were observed in a nontreated (control) group. Oral administration of LVA significantly inhibited acetic acid induced mucosal injury in both prophylactic and therapeutic groups in comparison with control rats (with mean scores of 4.2 and 4.2 vs 10.3, respectively; P < 0.05). Moreover, the data obtained showed that LVA therapy could significantly reduce serum and colonic mucosal IL-6, TNF-a, LTB4, and PGE2 levels in treated groups compared with nontreated controls. Figure 1 shows that LVA therapy is able to reduce significantly the serum level of TNF-a in prophylactic and therapeutic groups compared with control rats (P < 0.001); however, although there was a significant difference between prophylactic and control groups in the serum IL-6 level (P < 0.003), the difference between therapeutic and control groups was not significant (P < 0.058). Moreover, as shown in Fig. 2, LVA administration could significantly diminish the serum concentration of eicosanoids (LTB4 and PGE2) in prophylactic and therapeutic groups compared
232
A. Mirshafiey and B.H.A. Rehm
a 90 80
Serum IL-6 (Pg/ml)
70 60
∗
50 40 30 20 10 0 C
P
T
b 180 160
Serum TNFα (Pg/ml)
140 120
∗
100
∗
80 60 40 20 0 C
P
T
Fig. 1 Effects of low-viscosity sodium alginate (LVA) on serum levels of interleukin-6 (IL-6) (a) and tumor necrosis factor a (TNF-a) (b), 1 week after induction of colitis. C control rats, P prophylactic group rats, T treated rats. Each bar represents the mean ± the standard deviation (SD). Asterisks denote a significant difference between C and P and/or C and T groups, *P £ 0.05 (Mirshafiey et al. 2005a)
with the control group (P < 0.01). On the other hand, with use of the supernatant of homogenized colonic tissue samples, it was found that there was a significant difference between prophylactic and therapeutic groups compared with the control group in connection with the amounts of IL-6, TNF-a, LTB4, and PGE2 (P < 0.05).
Alginate and Its Comonomer Mannuronic Acid
a
233
7
Serum L TB4 (ng/ml)
6 5 4
∗
∗
3 2 1 0 C
P
T
b 14
Serum PGE2 (ng/ml)
12 10 8
∗
∗
P
T
6 4 2 0 C
Fig. 2 Effects of LVA on serum levels of leukotriene B4 (LTB4) (a) and prostaglandin E2 (PGE2) (b), 1 week after induction of colitis. C control rats, P prophylactic group rats, T treated rats. Each bar represents the mean ± S.D. Asterisks denote a significant difference between C and P and/or C and T groups, *P £ 0.05 (Mirshafiey et al. 2005a)
Figures 3 and 4 show the effect of a therapeutic protocol on the reduction of the production of cytokines (IL-6 and TNF-a) and eicosanoids (LTB4 and PGE2). Taken together, the results of this study demonstrate that LVA therapy palliates the progression of colonic inflammatory lesions in an experimental model of IBD. The
234
A. Mirshafiey and B.H.A. Rehm
a 160 140
∗
Tissue IL-6 (Pg/ml)
120
∗
100 80 60 40 20 0 C
P
T
b 350
Tissue TNFα (Pg/ml)
300 250
∗ 200
∗
150 100 50 0 C
P
T
Fig. 3 Amounts of IL-6 (a) and TNF-a (b) in colonic homogenate following 1 week of oral administration of LVA. C control rats, P prophylactic group rats, T treated rats. Each bar represents the mean ± S.D. Asterisks denote a significant difference between C and P and/or C and T groups, *P £ 0.05 (Mirshafiey et al. 2005a)
Alginate and Its Comonomer Mannuronic Acid
235
a 4500 4000
Tissue LTB4 (Pg/ml)
3500 3000 2500 2000
*
1500
*
1000 500 0 C
P
T
b 16000
Tissue PGE2 (Pg/ml)
14000 12000 10000
*
8000
*
6000 4000 2000 0 C
P
T
Fig. 4 Amounts of LTB4 (a) and PGE2 (b) in colonic homogenate following 1 week of oral administration of LVA. C Control rats P prophylactic group rats, T treated rats. Each bar represents the mean ± S.D. Asterisks denote a significant difference between C and P and/or C and T groups, *P £ 0.05 (Mirshafiey et al. 2005a)
beneficial effect of LVA is associated with a reduction of cytokine (IL-6 and TNF-a) and eicosanoid (LTB4 and PGE2) synthesis. Thus, LVA as a new therapeutic option can be recommended for IBD preclinical trials.
236
3
A. Mirshafiey and B.H.A. Rehm
Treatment of Experimental Chronic Ulcerative Colitis with Sodium Alginate
The purpose of this study was to test the therapeutic efficacy of sodium alginate in a rat model of trinitrobenzene sulfonic acid (TNBS) induced IBD (Razavi et al. 2008). This experiment was carried out using 77 Sprague-Dawley rats which were divided into five groups: normal, control, prophylactic, therapeutic, and experimental groups. The rats were killed 1, 2, 3, and 6 weeks after induction of colitis. The severity of colitis was graded macroscopically and assessed using serum and colonic mucosal cytokines and eicosanoids. Intrarectal TNBS (30 mg) produced a significant chronic ulcerative colitis, on the basis of the method of (Menozzi et al. 2006). The lesions were most severe on the seventh day after TNBS instillation, and then declined, but lesions were still observed after 6 weeks. The onset of oral LVA administration for prophylactic and therapeutic groups was 24 h before and after the induction of colitis, respectively. Macroscopic findings showed that intracolonic administration of TNBS resulted in acute (week 1) and chronic (week 6) IBD in control nontreated rats. The control rats developed colonic macroscopic damage such as diffuse hyperemia and ulcerations. Daily oral treatment with LVA (drinking 35.5 ± 2.8 ml of 0.5% W/V LVA solution per rat, which was equivalent 0.65–0.77 mg kg−1 LVA per day) significantly reduced the colonic damage score in prophylactic and therapeutic groups in weeks 1, 2, 3, and 6 after induction of colitis. TNBS administration also significantly enhanced the levels of serum and colonic mucosal cytokines (TNF-a and IL-6) and eicosanoids (LTB4 and PGE2), which paralleled the severity of colitis, whereas pretreatment (in the prophylactic group) and treatment with LVA were significantly able to reduce the colonic damage score, serum level, and colonic mucosal production of TNF-a, IL-6, LTB4, and PGE2 in prophylactic and therapeutic rat groups compared with nontreated controls. Figure 5a shows that LVA therapy was significantly able to reduce the serum level of TNF-a in prophylactic and therapeutic groups (during weeks 2, 3, and 6) compared with colitis induced in control rats (P < 0.05). Figure 5b represents a significant difference in the serum level of IL-6 in prophylactic and therapeutic groups (during weeks 1, 2, 3, and 6) compared with control rats (P < 0.05). Moreover, as shown in Fig. 6a, LVA administration could significantly diminish the serum concentration of LTB4 (during weeks 1, 3, and 6) in the prophylactic group and (during weeks 1 and 3) in the therapeutic group compared with the colitis-induced group (P < 0.05). Figure 6b shows that LVA therapy was significantly able to reduce the serum level of PGE2 in the prophylactic group (during week 6) and in the therapeutic group (during weeks 1 and 3) compared with colitis-induced rats (P < 0.05). In addition, with use of the supernatant of homogenized colonic tissue samples, it was found that there is a significant difference between prophylactic and therapeutic groups compared with the colitis-induced group in connection with the amounts of IL-6, TNF-a, and LTB4 (during weeks 1 and 6) (Figs. 7a, b, 8a), whereas there was no significant difference between prophylactic and therapeutic groups compared with the control group in colonic tissue PGE2 level following LVA therapy in an experimental model of colitis (Fig. 8b). Collectively, the data showed that LVA therapy is able to suppress chronic ulcerative colitis in an experimental model.
Alginate and Its Comonomer Mannuronic Acid
237
a 300 250
Pg / ml
200 T 150
P C
100 50 0 Week 1
Week 2
Week 3
Week 6
b 70 60
Pg / ml
50 T 40 P 30
C
20 10 0 Week 1
Week 2
Week 3
Week 6
Fig. 5 Effects of LVA on serum levels of TNF-a (a) and IL-6 (b) during weeks 1, 2, 3, and 6 after induction of colitis. C colitis-induced rats; P prophylactic group rats, T treated rats (Razavi et al. 2008)
4
Treatment of Experimental Immune Complex Glomerulonephritis by Sodium Alginate
The therapeutic efficacy of sodium alginate was studied in experimental immune complex glomerulonephritis (Mirshafiey et al. 2005). On the basis of the method of (Yamamoto et al. 1978), bovine serum albumin (BSA) nephritis as an
238
A. Mirshafiey and B.H.A. Rehm
a 3.5 3
Pg/ ml
2.5 T
2
P 1.5
C
1 0.5 0 Week 1
Week 2
Week 3
Week 6
b 12 10
Pg/ ml
8 T 6
P C
4 2 0 Week 1
Week 2
Week 3
Week 6
Fig. 6 Effects of LVA on serum levels of LTB4 (a) and PGE2 (b), 1 and 6 weeks after induction of colitis. C colitis-induced rats; P prophylactic group rats, T treated rats (Razavi et al. 2008)
experimental model of glomerulonephritis was induced in rats by a subcutaneous immunization and daily intravenous administration of BSA. Sodium alginate at two different doses (25 and 50 mg kg−1) was administered intraperitoneally to the treatment groups (T1 and T2) at regular 72-h intervals for 6 weeks. The onset of treatment was on day 42. Urinary protein was measured weekly and serum anti-BSA antibody was assessed by ELISA at different intervals. Rats were euthanized at the 12th experimental week and blood samples and kidney specimens were obtained. Blood urea nitrogen (BUN), serum creatinine, and serum cholesterol and triglyceride were measured at the time of death. Kidney specimens were processed for light and immunofluorescence microscopy examination.
Alginate and Its Comonomer Mannuronic Acid
239
a 200 180 160 Pg / ml
140
T
120 P
100
C
80 60 40 20 0 Week 1
Week 6
b 1400 1200 1000 Pg / ml
T 800
P
600
C
400 200 0 Week 1
Week 6
Fig. 7 Amounts of IL-6 (a) and TNF-a (b) in colonic tissue homogenate following 6 weeks of oral administration of LVA. C colitis-induced rats, P prophylactic group rats, T treated rats (Razavi et al. 2008)
The results of this experiment showed that treatment with sodium alginate could significantly reduce the urinary protein excretion and serum creatinine in treated rats compared with nontreated controls. Figure 9 shows the changes of the mean levels of urinary protein excretion between normal, patient, and treatment (T1 and T2) groups. The urinary protein excretion was significantly less at the end of the experiment (day 85) in LVA-treated rats (group T2, 50 mg kg−1) than in the nontreated controls. Comparison of BUN and serum creatinine concentrations among various groups showed only significant reduction (P < 0.05) in the level of serum creatinine in LVA-treated rats (T2 group) compared with
240
A. Mirshafiey and B.H.A. Rehm
a 4500 4000 3500
Pg/ml
3000
T
2500
P
2000
C
1500 1000 500 0 Week 1
Week 2
b 12000 10000
Pg/ ml
8000
T P
6000
C 4000 2000 0 Week 1
Week 2
Fig. 8 Amounts of LTB4 (a) and PGE2 (b) in colonic tissue homogenate following 6 weeks of oral administration of LVA. C colitis-induced rats, P prophylactic group rats, T treated rats (Razavi et al. 2008)
nontreated controls. Moreover, the anti-BSA antibody titers were significantly lower in LVA-treated rats (group T2, 50 mg kg−1) than in nontreated controls at the 12th week after immunization. The changes of the mean levels of anti-BSA antibody titers in the third, sixth, ninth, and 12th experimental weeks are illustrated in Fig. 10. There was no significant difference in the level of BUN and serum lipids (triglyceride and cholesterol) between different groups. The light microscopy examination of renal tissue revealed the severity of hypercellularity, glomerular infiltration of polymorphonuclear leukocytes (PMN), fibrinoid necrosis, and inter-
140 p T1 T2 N
Proteinuria (mg/day)
120 100 80 60 40 20 0 0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
Time (weeks)
Fig. 9 Time course of the mean values of proteinuria in different groups in immune complex glomerulonephritis: N normal rats (n = 10), P patient rats (n = 9), T1, T2 groups treated by intraperitoneal injections of LVA at two different doses, 25 and 50 mg kg−1, respectively (n = 2 × 9). The onset of intraperitoneal administration of LVA to T1 and T2 groups was on day 42. There were 14 intraperitoneal injections, the injection interval was 72 h, and the end of the therapeutic protocol was day 81. There was a significant difference between the patient group and the treated group (T2). P < 0.05 was considered significant (Mirshafiey et al. 2005b)
2 p T1 T2
A 492 nm
1.5
1
0.5
0 3
6
9
12
Time (weeks)
Fig. 10 Comparison of anti-bovine serum albumin (BSA) antibody titers in groups P, T1, and T2 during the course of acute serum sickness. Patient rats were divided into four subgroups (n = 5 + 5 + 5 + 9) and treated rats into T1 (n = 7 + 9) and T2 (n = 7 + 9) subgroups. Rats from patient subgroups were killed in four stages (third, sixth, ninth, and 12th weeks), whereas treated subgroups were killed in two stages (ninth and 12th weeks) after immunization to obtain their sera. Treated groups (T1 and T2) received intraperitoneal injections of LVA at two different doses, 25 and 50 mg kg−1, respectively. The onset of intraperitoneal administration of LVA to T1 and T2 groups was on day 42. As determined by the ELISA method, the highest titers occurred in the P group, the intermediate titers were in the T1 group, and the lowest titers were in the T2 group. At the 12th week, the reduction of anti-BSA antibody in the T2 group was significant compared with that in the P group. Values are means ± S.D (Mirshafiey et al. 2005b)
242
A. Mirshafiey and B.H.A. Rehm
stitial infiltration in various groups. Rats treated with LVA (group T2) showed a significant reduction in glomerular changes compared with nontreated controls. Immunofluorescence microscopy investigation of glomeruli revealed deposits of immune complexes in the mesangial areas and along the capillary walls of all the rats. Glomerular deposition of immune complex was significantly less intense in LVA-treated rats (group T2) than in nontreated controls. These findings suggest that treatment with sodium alginate as a new immunosuppressive agent can reduce proteinuria, and suppress the antibody production as well as the development of glomerular lesions in a rat model of immune complex glomerulonephritis.
5
Therapeutic Effects of M2000 on an Experimental Model of Rheumatoid Arthritis
This investigation was planned to explore the therapeutic potency of M2000, a novel designed nonsteroidal anti-inflammatory drug in an adjuvant-induced arthritis model (Mirshafiey et al. 2005). For this purpose, a production process for the commercially unavailable M2000 had to be established. An alginate epimerase negative Pseudomonas strain was engineered which synthesized homopolymeric polymannuronate. The polymannuronate was subjected to acid hydrolysis to obtain the monomer, which was further purified by gel filtration chromatography. On the basis of the method of (Cuzzocrea et al. 2005), arthritis was induced in Lewis rats by a single intradermal injection (0.1 ml) of heat-killed Mycobacterium tuberculosis (0.3 mg) in Freund’s incomplete adjuvant into the right foot pad. Fourteen days after injection of adjuvant, the contralateral left foot pad volume was measured. The rats with paw volumes 0.37 ml greater than those of normal paws were then randomized into treatment groups. Intraperitoneal administration of test drugs (M2000, 40 mg kg−1 per day to group T1 and indomethacin, 2 mg kg−1 per day to group T2) and oral administration of M2000 (40 mg kg−1 per day to group T3) were started on day 15 after adjuvant injection and continued until the final assessment on day 25. The left hind limb was removed for histological evaluation. The results showed that the oral administration as well as the intraperitoneal injection of M2000 into arthritic rats induced a significant reduction in paw edema. Histopathological assessment showed a reduced inflammatory cell infiltrate in joints of treated rats, and the number of osteoclasts present in the subchondral bone, tissue edema, and bone erosion in the paws were markedly reduced following M2000 therapy. The data in Fig. 11 demonstrate a time-dependent increase induced by adjuvant in the left hind paw volume (milliliters) of rats. The intraperitoneal injection of M2000 into arthritic rats could rapidly reverse paw edema as did indomethacin, and after 10 days of treatment paw swelling was significantly (P < 0.05) reduced in M2000-treated rats compared with vehicle-treated controls. On the other hand, as is apparent in Fig. 12, oral administration of M2000 (40 mg kg−1) in arthritic rats induced a significant reduction in paw edema. After 10 days of treatment, paw swelling was reduced in M2000-treated rats relative to the paw volume of vehicletreated rats. This difference was statistically significant. There was no macroscopic evidence of hind paw erythema and/or edema in the normal control rats.
4
Adjuvant Adjuvant + M2000
3.5
Adjuvant + INDO
PAW OEDEMA (ml)
3 2.5 2 1.5 1 0.5 0 0
4
8
12
15
20
22
24
25
Days
Fig. 11 Severity of arthritis score in treated groups (A + M) compared with nontreated control rats (a). A time-dependent increase induced by adjuvant in the left hind paw volume (milliliters) of rats is demonstrated. The intraperitoneal injection of b-d-mannuronic acid (M2000) to arthritic rats could rapidly reverse paw edema as did indomethacin, and after 10 days of treatment paw swelling was significantly (P < 0.05) reduced in M2000-treated animals compared with vehicletreated controls (Mirshafiey et al. 2005c) *P<0.001 vs Adjuvant 4 Adjuvant
3,5
PAW OEDEMA (ml)
Adjuvant + M2000
*
*
24
25
*
3
*
2,5 2 1,5 1 0,5 0 0
4
8
12
15
20
22
Days
Fig. 12 The oral administration of M2000 (40 mg kg−1) in arthritic rats induced a significant (P < 0.05) reduction in paw edema. After 10 days of treatment paw swelling was reduced in M2000-treated animals relative to the paw volume of vehicle-treated rats. This difference was statistically significant, P < 0.05 (Mirshafiey et al. 2005c)
244
A. Mirshafiey and B.H.A. Rehm 10 9
Histological damage score
8
VEHICLE M2000
∗
INDO
7 6 5 4 3 2 1 0 SHAM
AA
Fig. 13 Effect of M2000 treatment on histological damage score. The histological evaluation of the paws in the vehicle-treated arthritic rats reveals signs of severe arthritis along with inflammatory cell infiltrate. M2000 therapy could significantly reduce the pathological parameter compared with that of the control group. P < 0.05 was considered significant (Mirshafiey et al. 2005c)
As shown in Figs. 13 and 14, histological evaluation of the paws in the vehicletreated arthritic rats reveals signs of severe arthritis along with inflammatory cell infiltrate. Histopathological assessment showed a reduced inflammatory cell infiltrate in the joints of treated rats, and the number of osteoclasts present in the subchondral bone, tissue edema, and bone erosion in the paws were markedly reduced by both treatments, indicating that the drugs tested were effective in retarding synovial inflammation and prevented joint destruction. Treatment with M2000 and indomethacin resulted in preservation of hyaline in the articular cartilage. Moreover, subchondral bone was intact and the numbers of osteoclasts in the subchondral and trabecular bone space were greatly reduced. Taken together; the data obtained show that M2000, as a novel nonsteroidal anti-inflammatory drug, could be suggested as an anti-inflammatory drug for long-term administration.
6
Immunosuppressive Effect of M2000 in Experimental Multiple Sclerosis
The therapeutic potency of M2000 as a novel designed nonsteroidal anti-inflammatory drug with immunosuppressive property in a T-cell-mediated autoimmune disease was tested (Mirshafiey et al. 2005). The influence of M2000 on experimental autoimmune
Alginate and Its Comonomer Mannuronic Acid
245
Fig. 14 Representative histopathological slides of a hind limb joint of a healthy Lewis rat (a), a rat with adjuvant-induced arthritis (b), and an arthritic rat treated with M2000 (c). The joint of the rat which had been treated with M2000 (c) shows significantly fewer signs of joint destruction (Mirshafiey et al. 2005c)
encephalomyelitis (EAE), an animal model of multiple sclerosis, induced by myelin basic protein (MBP) was assessed. M2000 at two doses, 40 and 80 mg kg−1 per day, was administered intraperitoneally to prevention and treatment groups, respectively. The onset of intraperitoneal injections of M2000 for prophylactic and therapeutic groups was on the first and seventh days after immunization. Rats were divided at random into a prevention group with two subgroups: M1 – M2000-pretreated rats, which received in total 18 intraperitoneal injections from 1 day before immunization (day 1) to day 16 after immunization; C1 – control patient rats, which received intraperitoneal injections of saline during the same period (from day 1 to day 16). The treatment group included two subgroups: M2 – M2000-treated rats, which received in total 14 intraperitoneal injections from day 7 to day 20 after immunization; C2 – control patient rats, which received intraperitoneal injections of saline during the same period. Rats were killed on day 21 after immunization. The results of this experiment showed that the treatment of EAE with M2000 could significantly suppress disease development both prophylactically and therapeutically; the onset and symptoms of EAE in Lewis rats could be suppressed following the administration of M2000. Clinical improvement was accompanied by a marked decrease in the mean numbers of vessels with perivascular cellular
246
A. Mirshafiey and B.H.A. Rehm
infiltration in M2000-treated rats compared with nontreated control. Disease suppression was associated with a marked suppression of MBP-specific T-cell reactivity in vitro, without any evidence for a generalized impairment of T-cell activity. Following the chronological changes of the clinical score of each group of the EAE rats (Figs. 15, 16), the first clinical signs appeared in some rats of each group on day 10 and then the symptoms reached their maximum level on day 13 after immunization. The maximum severity score of EAE was 5 in each group. The rats pretreated and treated with M2000 were less affected than the controls (Figs. 15 and 16). Three rats from the M1 subgroup did not develop apparent signs of EAE during the period of observation, whereas all of the rats from the C1 and C2 subgroups developed typical signs (clinical score more than 4) of EAE. As shown in Figs. 15 and 16, the mean clinical scores of rats for subgroups M1 and M2 were significantly less than those of control rats (the two-way repeated measures analysis of variance, P < 0.05). The mortality rate in the M1 subgroup (40 mg kg−1 per day) was 0/7 compared with 0/8 for the C1 subgroup, whereas, interestingly, the mortality rate in the M2 subgroup (80 mg kg−1 per day) was 0/6 compared with 2/8 in the C2 subgroup. The mean body weight more reduced in controls than in subgroup M1 rats after day 12 (Fig. 17). Moreover, lymph node cells from the rats treated with M2000 exhibited a reduced proliferative response to MBP in contrast to those of control patient rats, whereas the proliferative responses of lymph node cells to Con A were almost the same in both groups (Fig. 18). On the other hand, the histopathological assessment of brain and spinal cord sections in treated and nontreated groups showed that the mean numbers of vessels with perivascular cellular infiltration in brain (0.57 ± 0.79) and spinal cord (2.0 ± 2.65) in M2000-treated rats were
Mean clinical score
Treatment 5 4.5 4 3.5 3 2.5 2 1.5 1 0.5 0
Saline M2000(40)
*
*
* *
*
*
*
* *
0
1
2
3
4
5
6
7
8
9 10 11 12 13 14 15 16 17 18 19 20 21 22
Days post immunization
Fig. 15 Amelioration of myelin basic protein (MBP) induced experimental autoimmune encephalomyelitis (EAE) by preventive application of M2000. Fifteen Lewis rats were immunized with 50 mg MBP in complete Freund’s adjuvant into one hind footpad and were then divided into two subgroups. M2000-pretreated rats (n = 7) received in total 18 intraperitoneal injections of M2000 (40 mg kg−1 per day) from 1 day before immunization to day 16 after immunization. Control rats (n = 8) received saline intraperitoneal injections during the same period. M2000 therapy caused a significant reduction in clinical score compared with that for nontreated controls (P < 0.05). Bars indicate ± SD (Mirshafiey et al. 2005d)
Alginate and Its Comonomer Mannuronic Acid
247 Treatment
mean clinical score
4 3.5 3 2.5 2 1.5
Saline M2000(80)
1 0.5
∗
0 0
1
3
2
4
5
6
7
8
∗
∗
∗
9 10 11 12 13 14 15 16 17 18 19 20 21 22
Days post immunization
Fig. 16 Amelioration of MBP-induced EAE by therapeutic application of M2000. Lewis rats were immunized with 50 mg MBP in CFA into one hind footpad and were then divided into two subgroups. M2000-treated rats (n = 6) received in total 14 intraperitoneal injections of M2000 (80 mg kg−1 per day) from day 7 to day 20 after immunization. Control rats (n = 8) received intraperitoneal injections of saline during the same period. M2000 therapy caused a significant reduction in clinical score compared with that for nontreated control (P < 0.05) (Mirshafiey et al. 2005d) 180 Saline M2000
170
Mean weight
160 150 140 130 120 110 100 0
2
4
6
8
10
12
14
16
18
20
22
Days post immunization
Fig. 17 The higher weight gain in M2000-treated rats compared with nontreated controls during the EAE period. M2000-treated rats (n = 7) received in total 18 intraperitoneal injections of M2000 (40 mg kg−1 per day) from 1 day before immunization to day 16 after immunization. Control rats (n = 8) received intraperitoneal injections of saline during the same period. Bars indicate SD (Mirshafiey et al. 2005d)
significantly less than in nontreated controls, with mean values of 2.43 ± 3.41 for brain and 10.71 ± 8.76 for spinal cord, respectively. Collectively, our data suggest that M2000 may provide a novel therapeutic option for T-cell-mediated autoimmune diseases in humans.
248
A. Mirshafiey and B.H.A. Rehm 0.4 0.35
Optical Density
0.3 medium
0.25
MBP5 MBP10 ConA
0.2 0.15 0.1 0.05 0 Saline
M2000
Fig. 18 The effect of M2000 at two different doses (5 and 10 mg ml−1) on proliferative responses of regional lymph node cells from EAE rats, “n = 4, for each group, on day 21 after immunization” (Mirshafiey et al. 2005d)
7 Treatment of Experimental Nephrotic Syndrome with M2000 The therapeutic effect of the M2000 molecule was tested in adriamycin-induced nephropathy (Mirshafiey et al. 2004a). To induce the experimental nephrotic syndrome, adriamycin was given once by a single intravenous injection (7.5 mg kg−1) through the tail vein. Six days after injection of adriamycin, a therapeutic protocol was developed by intraperitoneal administration of 30 mg kg−1 M2000 solution to the treatment 1 (T1) group and intraperitoneal injection of piroxicam (0.3 mg kg−1) for the treatment 2 (T2) group. In total there were 14 intraperitoneal injections, of which five injections were performed day after day and nine injections were carried out at regular 48-h intervals. The therapeutic protocol was terminated on day 28 and the rats were killed on day 43. The treated patient rats showed a significant reduction in proteinuria, BUN, serum creatinine, and serum cholesterol, and administration of M2000 could significantly diminish the serum level of IL-6 in treated rats compared with nontreated controls. Moreover, treatment with M2000 significantly reduced the number of glomerular leukocytes, hypercellularity, and hydropic change in the capillary network within the renal cortex and decreased tubular casts. The changes of the mean levels of urinary protein excretion between normal, patient, T1, and T2 groups are shown in Fig. 19. This experiment showed that intraperitoneal administration of M2000 (30 mg kg−1) could exert its therapeutic effects on adriamycin-induced nephropathy. For exact evaluation of the effects of M2000, the experiment was terminated 14 days after the last injection of M2000. The urinary protein excretion was significantly less in M2000-treated rats than in nontreated controls and piroxicam-treated rats. Figure 20 shows the antiproteinuric effect of M2000 therapy compared with that of piroxicam at the end of experiment (day 42). Here there was significant difference
Alginate and Its Comonomer Mannuronic Acid
249
1000 900 urine protein mg/24
800 700 600 500 400 300 200
N
P
100
T1
T2
0 0
5
14
21
28
35
42
DAYS
Fig. 19 Time course of the mean values of proteinuria in different groups in experimental nephrosis: N normal rats (n = 8), P patient (nontreated) rats that show an increase of proteinuria levels after day 5 (n = 9), T1, T2 groups treated by intraperitoneal injections of M2000 (30 mg kg−1) and piroxicam (0.3 mg kg−1), respectively (n = 2 × 8). Note: a single intravenous injection of adriamycin (7.5 mg kg−1 body weight) induced a severe nephrotic syndrome. The onset of intraperitoneal administration of M2000 and piroxicam to groups T1 and T2 was on day 6 (after development of disease).There were significant differences between nontreated (P) and M2000-treated rats (T1). P < 0.05 was considered significant (Mirshafiey et al. 2004a) 1200
urine protein mzg / 24
1000
800
600
400
200
0 N
P
T1
T2
Fig. 20 Effect of M2000 on proteinuria at the end of the experiment (day 42) in different groups: N normal rats (8), P patients rats (9), T1 rats treated with M2000 (8), T2 rats treated with piroxicam (8). Each bar represents the mean ± SD. The comparison of the antiproteinuric effect of M2000 between groups P and T1 showed a significant difference, “P < 0.05 (Mirshafiey et al. 2004a)
250
a
A. Mirshafiey and B.H.A. Rehm
45 40 35
BUN mg/ dl
30 25 20 15 10 5 0
b
N
P
T1
T2
N
P
T1
T2
1.6
Serum Creatinine mg /dl
1.4 1.2 1 0.8 0.6 0.4 0.2 0
Fig. 21 Effect of M2000 on serum blood urea nitrogen and serum creatinine in different groups: N normal rats (8), P patient rats (9), T1 rats treated with M2000 (8), T2 rats treated with piroxicam (8). Each bar represents the mean ± SD. a BUN concentration in groups N, P, T1, and T2. BUN concentration in group T1 vs. group P was significant, P < 0.05. b Creatinine concentration in groups N, P, T1, and T2. Creatinine concentration in group T1 vs. group P was significant, P < 0.05 (Mirshafiey et al. 2004a)
between T1 and patient groups. In Fig. 21a, the amounts of BUN are compared between the different groups (normal, patient, T1, and T2). This figure shows that the difference between treated rats (T1 group) and nontreated rats (patient group)
Alginate and Its Comonomer Mannuronic Acid
251
is significant. In Fig. 21b, the comparison of serum creatinine concentration among various groups shows that there is a significant difference between T1 and patient groups, whereas there were no significant differences in the levels of urinary urea nitrogen and urine creatinine concentration between M2000-treated rats and nontreated controls (data not shown). In addition, serum cholesterol and triglyceride levels were significantly elevated in nephrotic rats when compared with the healthy controls at the end of the experiment. Intraperitoneal injections of M2000 solution into patient rats significantly reduced serum cholesterol levels in nephrotic rats (Fig. 22). Figure 23 shows the effect of M2000 therapy on the reduction of IL-6 production in treated patient rats (group T1). The difference between T1 and patient groups in terms of IL-6 concentration was significant. Light microscopy examination of renal tissue revealed the severity of hypercellularity, glomerular infiltration of PMN, hydropic change in the capillary network within the renal cortex, and the existence of tubular casts in the various groups. The rats treated with M2000 showed a significant reduction in glomerular changes compared with nontreated controls (Table 1). These data suggest that M2000 therapy can ameliorate proteinuria and suppress the progression of glomerular lesions in an experimental model of nephrosis.
350
serum cholesterol mg / dl
300
250
200
150
100
50
0 N
P
T1
T2
Fig. 22 Effect of M2000 on serum cholesterol level in different groups: N normal rats (8), P patient rats (9), T1 rats treated with M2000 (8), T2 rats treated with piroxicam (8). Each bar represents the mean ± SD. Cholesterol concentration in group T1 vs. group P was significant, P < 0.05 (Mirshafiey et al. 2004a)
252
A. Mirshafiey and B.H.A. Rehm 45
serum IL-6 pg/ ml
40 35 30 25 20 15 10 5 0 P
T1
T2
Fig. 23 Effect of M2000 on serum IL-6 level in different groups: N normal rats (8), P patient rats (9), T1 rats treated with M2000 (8), T2 rats treated with piroxicam (8). Each bar represents the mean ± SD. The Difference between group T1 and group P in terms of IL-6 concentration was significant, P< 0.05, (Mirshafiey et al. 2004a)
t.1
Table 1 Light microscopy findings of kidney histological lesions in various groups
t.2 t.3
Polymorphonuclear Group No. of rats Hypercellularity leukocyte infiltration
Cast
Hydropic change
t.4 t.5 t.6 t.7 t.8
N P T1 T2 H
– 2.14 ± 0.69 1.14 ± 0.63 1.87 ± 1.1 –
– 1±0 – –
t.9 t.10 t.11 t.12 t.13
8 9 8 8 8
0.40 ± 0.53 1 ± 0.57 0.57 ± 0.53 0.71 ± 0.46 0.42 ± 0.53
– 1.71 ± 0.48 1 1.6 ± 0.76 –
–
Values are expressed as the mean ± the standard deviation. Semiquantitative scoring of histological lesions shows the significant differences between groups P and T1, in hypercellularity, in glomerular polymorphonuclear leukocyte infiltration, hydropic change in the capillary network within the renal cortex, and in tubular casts, (P < 0.05). Group H was the healthy control receiving b-d-mannuronic acid (M2000; 30 mg kg−1)
8
Therapeutic Effects of M2000 in Experimental Immune Complex Glomerulonephritis
The therapeutic efficacy of the novel anti-inflammatory agent M2000 in an experimental model of immune complex glomerulonephritis was evaluated (Mirshafiey et al. 2007). BSA nephritis as an experimental model of glomerulonephritis was induced in rats by a subcutaneous immunization and daily intravenous administration of BSA. Rats were divided randomly into five groups. normal; patient; patient groups treated with M2000 and piroxicam (T1 and T2, respectively); a healthy control receiving M2000. The M2000 solution (30 mg kg−1) was administered intraperitoneally at regular 48-h intervals for 4 weeks. The onset of treatment was on day 56. Urinary protein was measured weekly and serum anti-BSA antibody was assessed by ELISA at different intervals. The rats were killed on day 84 and blood samples and
Alginate and Its Comonomer Mannuronic Acid
253
kidney specimens were obtained. Serum (creatinine, BUN, cholesterol, and triglyceride) and urine (protein, urea, and creatinine) determinants were measured at the time of death. Kidney specimens were processed for light and immunofluorescence microscopy examination. The results showed that M2000 therapy could significantly reduce the urinary protein excretion in treated rats compared with nontreated controls. The anti-BSA antibody titer was lower in treated rats than in controls at the 12th experimental week. There were no significant differences in the levels of serum determinants, urine urea, and creatinine between control and treated rats. PMN infiltration and glomerular immune complex deposition were less intense in treated rats than in controls. Figure 24 shows the changes of the mean levels of urinary protein excretion between various groups; normal, patient, T1, and T2. The urinary protein excretion was significantly less in M2000-treated rats than in nontreated controls. Figure 25 shows the antiproteinuric effect of M2000 therapy compared with patient and piroxicam (T2) groups at the end of the experiment (day 84). Here, the reduction of proteinuria in T1 compared with patient and T2 groups was significant. Figure 26 illustrates the changes of the mean levels of anti-BSA antibody titers in the sixth, ninth, and 12th experimental weeks. The anti-BSA antibody titers were significantly lower in M2000-treated rats (T1 group) than in piroxicamtreated rats (T2 group) and nontreated controls (patient group) at the end of the experiment (day 84).
Fig. 24 Time course of the mean values of proteinuria in different groups in immune complex glomerulonephritis: N normal rats (n = 8), P patient rats (n = 9), T1, T2 rats treated with intraperitoneal injections of M2000 (30 mg kg−1) and piroxicam (0.3 mg kg−1), respectively (n = 2 × 8). The onset of intraperitoneal administration of M2000 and piroxicam to T1 and T2 groups was on day 56. There were 15 intraperitoneal injections, the injection interval was 48 h, and the end of the therapeutic protocol was day 84. There was a significant difference between the M2000-treated group (T1) and the patient group. P < 0.05 was considered significant (Mirshafiey et al. 2007)
254
A. Mirshafiey and B.H.A. Rehm 250
proteinuria (mg/ day)
200
150
100
50
0 N
P
T1
T2
Fig. 25 Comparison of the antiproteinuric effect of M2000 between different groups at the end of experiment (week 12). N normal rats (8), P patient rats (9), T1, T2 rats treated with M2000 (30 mg kg−1) and piroxicam (0.3 mg kg−1), respectively (2 × 8 = 16). Each bar represents the mean ± S.D. The antiproteinuric effect for group T1 vs. group P was significant, P < 0.05 (Mirshafiey et al. 2007)
1.8 1.6 1.4
A 492 nm
1.2
P
1 T1 0.8 T2
0.6 0.4 0.2 0 6
9
12
Time (weeks)
Fig. 26 Anti-BSA antibody titers in groups P, T1, and T2 during the course of acute BSA nephritis, as determined by an ELISA method. The highest titer occurs in group P, the intermediate in group T2, and the lowest in group T1. Values are means ± S.D. The difference between the M2000-treated group (T1) and the patient group (P) at the 12th week was significant. P < 0.05 was considered significant (Mirshafiey et al. 2007)
Alginate and Its Comonomer Mannuronic Acid
255
Light microscopy examination of renal tissue revealed the severity of glomerular infiltration of PMN in nontreated rats (patient group) and in rats treated with piroxicam (T2 group) compared with M2000-treated rats (T1 group) (data not shown). Immunofluorescence microscopy investigation of glomeruli revealed that glomerular immune complex deposition was less intense in rats treated with M2000 (T1 group) than in controls (patient group) and in rats treated with piroxicam (data not shown). These findings suggested that treatment with M2000 (C6H10O7) can reduce proteinuria, diminish antibody production, and suppress the progression of disease in a rat model of immune complex glomerulonephritis.
9
Tolerability and Pharmacotoxicology of M2000
The pharmacotoxicology study of M2000 was carried out on animal models based on the evaluation of serum and urine determinants, structure of kidney, gastrointestinal tolerability, and body temperature (Mirshafiey et al. 2004). Moreover, the WEHI-164 (fibrosarcoma) cell line was used to assay tolerability and matrix metalloproteinase type 2 (MMP-2) activity, on the basis of the method of (Heussen and Dowdle 1980). MMP-2 activity was assessed using zymography. Toxicology studies were categorized on the basis of: 1. Serum and urine determinants: To evaluate the side effects of M2000, assessment of kidney function was performed in healthy rats receiving the test drug after 12 intraperitoneal injections of M2000 (30 mg kg−1 each 48 h) and was based on the measurements of serum creatinine, BUN, urinary protein excretion, urine urea, as well as the plasma concentration of triglyceride and cholesterol. 2. Histological examinations: Renal tissues were assessed using light microscopy. Glomerular lesions were graded on a scale of 0–3 (0, negative; 1, mild; 2, moderate; 3, marked) according to four parameters: hypercellularity, glomerular infiltration of PMN, hydropic change in the capillary network within the renal cortex, and the presence of tubular casts. 3. Gastroulcerogenic study: After 12 intraperitoneal injections of M2000 (30 mg−1 kg−1 each 48 h) for the group of healthy control rats, as well as oral administration of M2000 (40 mg−1kg−1 per day) for 10 days to group T3, the rats were killed and the stomach and the duodenum were dissected out. Scoring was done on the basis of at least one gastric ulcer or one hemorrhagic erosion. 4. Body temperature: The course of rectal temperature of rats which received 12 intraperitoneal injections of M2000 solution (30 mg−1 kg−1 each 48 ) was monitored continuously by thermoelements and compared with the baseline temperature in normothermic controls. Cytotoxicity analysis showed a much higher tolerability for M2000 than for other drugs tested (diclofenac, piroxicam, and dexamethasone). The inhibitory effect of M2000 on MMP-2 activity was significantly greater than that of dexsamethasone
256
A. Mirshafiey and B.H.A. Rehm
and of piroxicam at a concentration of 200 mg ml−1. Moreover, data analysis showed no significant differences in the levels of serum determinants (BUN, creatinine, cholesterol, and triglyceride) and urine determinants (protein excretion and creatinine) between the normal group and the healthy experimental group challenged with M2000 (Table 2), whereas a significant increase of urine urea in the healthy group compared with the normal group revealed the advantage using M2000 in healthy controls (Table 2). The histological study of kidney specimens obtained from the normal group and healthy controls receiving M2000 showed no glomerular changes Table 2 Toxicological inspection of M2000 administration on serum and urine determinants
No. of Group rats
Urine Blood urea Serum Serum Serum protein nitrogen creatinine cholesterol triglyceride (mg (mg dl−1) (mg dl−1) (mg dl−1) (mg dl−1) day−1)
Urine urea
N H
16 ± 2 17 ± 4
939 ± 300 1,284 ± 226
8 8
0.7 ± 0.1 0.8 ± 0.1
59 ± 3 56 ± 6
79 ± 9 89 ± 23
(mg dl−1)
8±3 6±2
The healthy group receiving M2000 (H) versus the normal (N) group shows no significant difference in serum and urine determinants.
Piroxicam 25000
Dexamethasone M2000 Diclofenac
Number of Cells
20000
15000
10000
5000
0 0
10
20
40
80
100
150
200
Concentration (ug /mI)
Fig. 27 Cytotoxic effect of M2000. Proliferative response of fibrosarcoma (WEHI-164) cell line to M2000 at different doses (10–200 mg ml−1) compared with those of diclofenac, piroxicam, and dexamethasone. LD50 for diclofenac, piroxicam, and dexamethasone was 25, 80, and 80 mg ml−1, respectively. In contrast, WEHI-164 as a sensitive cell line showedhigh tolerability against increasing amounts of M2000 (Mirshafiey et al. 2004b)
Alginate and Its Comonomer Mannuronic Acid
257
in the healthy group in comparison with the normal group (data not shown). On the other hand, administration of M2000 had no influence on body temperature of normothermic rats which had received the test drug. Additionally, the results of this experiment showed that 12 intraperitoneal injections of M2000 solution (30 mg kg−1 each 48 h) as well as oral administration of M2000 (40 mg kg−1 per day) for 10 days could not provoke gastromucosal lesions in the rats of the experimental group. Figure 27 shows the proliferative response of a fibrosarcoma (WEHI-164) cell line to M2000 at different doses (10–200 mg ml−1) compared with the proliferative responses to diclofenac, piroxicam, and dexamethasone. The tolerability and biocompatibility of WEHI-164, as a sensitive cell line against increasing amounts of M2000, was very high, whereas 50% of cells died when diclofenac, dexamethasone, and piroxicam were added to tissue culture at doses of 25, 80, and 80 mg ml−1, respectively. M2000 showed no cytotoxic effect compared with steroidal and nonsteroidal drugs tested. Figure 28 presents the dose response analysis of the effect of M2000 on MMP-2 activity compared with the effect of various drugs. The inhibitory effect of M2000 at 1.2
Diclofenac
Relative Expression of MMP Activity
Piroxicam Dexamethasone
1
M2000 0.8
0.6
0.4
0.2
0 0
10
20
40
80
100
150
200
Concentration(ug/ ml)
Fig. 28 The inhibitory effect of M2000 on matrix metalloproteinase type 2 activity. Fibrosarcoma cell lines (2 × 104 cell wall−1) were incubated overnight with increasing doses of M2000. Diclofenac-, piroxicam- and dexamethasone-treated cells were used as controls. M2000-treated and nontreated cells were investigated in triplicate. The inhibitory activity of M2000 paralleledthat of diclofenac at doses of 10, 20, 40, and 200 mg ml−1, whereas the inhibitory effect of M2000 at concentrations of 20, 40, and 80 mg ml−1 was significantly more than that of dexamethasone (P < 0.05). Moreover, the inhibitory activity of this novel agent at a dose of 200 mg ml−1 was significantly more than that of dexamethasone and that of piroxicam, P < 0.05 (Mirshafiey et al. 2004b)
258
A. Mirshafiey and B.H.A. Rehm
concentrations of 20, 40, 80, and 200 mg ml−1 was significantly more than that of dexamethasone (P < 0.05), and this difference was significant between M2000 and piroxicam at a concentration of 200 mg ml−1, (P < 0.05).In contrast, the inhibitory activity of this novel agent paralleled that of diclofenac at doses of 10, 20, 40, and 200 mg ml−1.
10
Outlook
Alginates have been considered for medical applications such as drug delivery, wound coverage material, and cell encapsulation/transplantation. However, only recently have alginate and its comonomer M2000 been described as potential drugs suitable, in particular, for the treatment of inflammatory diseases. As compiled in this chapter, alginate as well as its comonomer M2000 show efficacy as antiinflammatory drugs when tested in various inflammatory diseases. The direct application of alginate/M2000 as a medical drug requires the production of alginates with consistent composition and high purity as well as the production of tailor-made alginates such as homopolymeric polymannuronate. Hence, the use of (engineered) microorganisms as production organisms is becoming increasingly attractive and fermentative production might in future be the only feasible production process.
References Ågren MS (1996) Four alginate dressings in the treatment of partial thickness wounds: a comparative experimental study. Br J Plastic Surg 49:129–134 Asada M, Sugie M, Inoue M, Nakagomi K, Hongo S, Murata K, Irie S, Takeuchi T, Tomizuka N, Oka S (1997) Inhibitory effect of alginic acids on hyaluronidase an on histamine release from mast cells. Biosci Biotechnol Biochem 61:1030–1032 Bucke C (1987) Cell immobilization in calcium alginate. Methods Enzymol 135:175–188 Cuzzocrea S, Mazzon E, Paola RD, Genovese T, Muià C, Caputi AP, Salvemini D (2005) Effects of combination M40403 and dexamethasone therapy on joint disease in a rat model of collagen-induced arthritis. Arthritis Rheum 52:1929–1940 Efentakis M, Buckton G (2002) The effect of erosion and swelling on the dissolution of theophylline from low and high viscosity sodium alginate matrices. Pharm Dev Technol 7:69–77 Gong YF, Huang ZJ, Qiang MY, Lan FX, Bai GA, Mao YX, Ma XP, Zhang FG (1991) Suppression of radioactive strontium absorption by sodium alginate in animals and human subjects. Biomed Environ Sci 4:273–282 Hasegawa T, Takahashi T, Inada Y, Yamada C, Tanaka Y (1989) Reparative effects of sodium alginate (Alloid G) on radiation stomatitis. Nippon Igaku Hoshasen Gakkai Zasshi 25:1047–1051 Heussen C, Dowdle EB (1980) Electrophoretic analysis of plasminogen activator in polyacrylamide gels containing sodium dodecyl sulfate and copolymerized substrates. Anal Biochem 102:196–202 Hori Y, Winans AM, Huang CC, Horrigan EM, Irvine DJ (2008) Injectable dendritic cell-carrying alginate gels for immunization and immunotherapy. Biomaterials 29:3671–3682 Jeong HJ, Lee SA, Moon PD, Na HJ, Park RK, Um JY, Kim HM, Hong SH (2006) Alginic acid has anti-anaphylactic effects and inhibits inflammatory cytokine expression via suppression of nuclear factor-kappaB activation. Clin Exp Allergy 36:785–794
Alginate and Its Comonomer Mannuronic Acid
259
Kammerlander G, Eberlein T (2003) An assessment of the wound healing properties of Algisite M dressings. Nurs Times 99:54–56 Lee JH, Kim WG, Kim SS, Lee JH, Lee HB (1997) Development and characterization of an alginate-impregnated polyester vascular graft. J Biomed Mater Res 36:200–208 Li HB, Jiang H, Wang CY, Duan CM, Ye Y, Su XP, Kong QX, Wu JF, Guo XM (2006) Comparison of two types of alginate microcapsules on stability and biocompatibility in vitro and in vivo. Biomed Mater 1:42–47 McGlashan JA, Johnstone LM, Sykes J, Strugala V, Dettmar PW (2009) The value of a liquid alginate suspension (Gaviscon Advance) in the management of laryngopharyngeal reflux. Eur Arch Otorhinolaryngol 266:243–251 Menozzi A, Pozzoli C, Poli E, Lazzaretti M, Grandi D, Coruzzi G (2006) Long-term study of TNBS-induced colitis in rats: focus on mast cells. Inflamm Res 55:416–422 Mirshafiey A, Rehm BH, Sahmani AA, Naji A, Razavi A (2004a) M-2000, as a new anti-inflammatory molecule in treatment of experimental nephrosis. Immunopharmacol Immunotoxicol 26:611–619 Mirshafiey A, Khorramizadeh MR, Saadat F, Rehm BH (2004b) Chemopreventive effect of M2000, a new anti-inflammatory agent. Med Sci Monit 10:PI105–PI109 Mirshafiey A, Cuzzocrea S, Rehm B, Mazzon E, Saadat F, Sotoude M (2005a) Sodium alginate as a novel therapeutic option in experimental colitis. Scand J Immunol 61:435–441 Mirshafiey A, Borzooy Z, Abhari RS, Razavi A, Tavangar M, Rehm BH (2005b) Treatment of experimental immune complex glomerulonephritis by sodium alginate. Vascul Pharmacol 43:30–35 Mirshafiey A, Cuzzocrea S, Rehm B, Mazzon E, Saadat F, Sotoude M (2005c) Treatment of experimental arthritis with M2000, a novel designed non-steroidal anti-inflammatory drug. Scand J Immunol 61:435–441 Mirshafiey A, Matsuo H, Nakane S, Rehm BH, Koh CS, Miyoshi S (2005d) Novel immunosuppressive therapy by M2000 in experimental multiple sclerosis. Immunopharmacol Immunotoxicol 27:255–265 Mirshafiey A, Rehm B, Sotoude M, Razavi A, Abhari RS, Borzooy Z (2007) Therapeutic approach by a novel designed anti-inflammatory drug, M2000, in experimental immune complex glomerulonephritis. Immunopharmacol Immunotoxicol 29:49–61 Morris GP, Beck PL, Herridge MS, Depew WT, Szewczuk MR, Wallace JL (1989) Hapteninduced model of chronic inflammation and ulceration in the rat colon. Gastroenterology 96: 795–803 Murata Y, Jinno D, Liu D, Isobe T, Kofuji K, Kawashima S (2007) The drug release profile from calcium-induced alginate gel beads coated with an alginate hydrolysate. Molecules 12: 2559–2566 Odell EW, Oades P, Lombardi T (1994) Symptomatic foreign body reaction to haemostatic alginate. Br J Oral Maxillofac Surg 31:165–169 Ohsumi H, Hirata H, Nagakura T, Tsujii M, Sugimoto T, Miyamoto K, Horiuchi T, Nagao M, Nakashima T, Uchida A (2005) Enhancement of perineurial repair and inhibition of nerve adhesion by viscous injectable pure alginate sol. Plast Reconstr Surg 116:823–830 Oshitani T, Okada K, Kushima T, Suematsu T, Obayashi K, Hirata Y, Takada Y, Ishida T, Yoshida M, Narabayashi I, et al. (1990) Clinical evaluation of sodium alginate on oral mucositis associated with radiotherapy. Nippon Gan Chiryo Gakkai Shi 20:1129–1134 Razavi A, Khodadadi A, Eslami MB, Eshraghi S, Mirshafiey A (2008) Therapeutic effect of sodium alginate in experimental chronic ulcerative colitis. Iran J Allergy Asthma Immunol 7:13–18 Rehm BHA (1998) Alginate lyase from pseudomonas aeruginosa CF1/M1 prefers the hexameric oligomannuronate as substrate. FEMS Microbiol Lett 165:175–180 Rehm BH, Valla S (1997) Bacterial alginates: biosynthesis and applications. Appl Microbiol Biotechnol 48:281–288 Selmi TA, Verdonk P, Chambat P, Dubrana F, Potel JF, Barnouin L, Neyret P (2008) Autologous chondrocyte implantation in a novel alginate-agarose hydrogel: outcome at two years. J Bone Joint Surg Br 90:597–604
260
A. Mirshafiey and B.H.A. Rehm
Sevgi F, Kaynarsoy B, Ertan G (2008) An anti-inflammatory drug (mefenamic acid) incorporated in biodegradable alginate beads: development and optimization of the process using factorial design. Pharm Dev Technol 13:5–13 Skjak-Braek G, Grasdalen H, Larsen B (1986) Monomer sequence and acetylation pattern in some bacterial alginates. Carbohydr Res 15:239–250 Siebers U, Horcher A, Bretzel RG, Federlin K, Zekorn T (1997) Alginate-based microcapsules for immunoprotected islet transplantation. Ann N Y Acad Sci 831:304–312 Yamada T, Marshall S, Specian RD, Grisham MB (1991) A Comparative analysis of two models of colitis in rats. Gastroenterology 102:1524–1534 Yamamoto T, Kihara I, Morita T, Oite T (1978) Bovine serum albumin (BSA) nephriti In rats: I. Experimental model. Acta Pathol Jpn 28:859–866 Yang D, Jones KS (2008) Effect of alginate on innate immune activation of macrophages. J Biomed Mater Res A (in press)
Index
A Absorbent dressings, 212 Absorption, 221 Acetic acid, 215 Acetophenone, 191 Acetyl-CoA, 55 Acetyl groups, 10 Acidovorax facilis, 188 Acid precipitation, 99 Acinetobacter calcoaceticus, 195 Agar, 183 Agarose, 200 Agriculture, 90 Agrobacterium tumefaciens, 87 Air lift reactors, 168 Alachlor, 198 Alcaligenes faecalis, 191 Alcohol dehydrogenase, 185 Alcohols, 215 Alg8, 62 Alg44, 55 AlgA, 60 AlgB, 117, 121 AlgC, 60 algC, 118, 140 AlgD, 60 algD, 140 algD operon, 118 algD promoter, 118 AlgE, 55, 65, 139 AlgE1, 8 AlgE4, 8, 105 AlgE5, 8 AlgE6, 8 AlgE7, 8 algF, 109 algG, 62 AlgG, 55, 102 algI, 109
Alginate, 2, 164 acetylation, 109 biosynthesis, 55 biosynthesis gene cluster, 58 gene regulation, 117 lyase, 62 polymerase, 56 polymerization, 63 production, 167 sequencing, 108 sodium alginate, 182 uptake, 74 Alginate lyases, 85, 108 Alginic acid, 2 algJ, 109 AlgJ, 109 AlgK, 55 algk, 62 AlgL, 55, 62 AlgM1, 83 AlgM2, 83 AlgQ1, 81 AlgQ2, 81 algR, 119 AlgR, 117 AlgS, 83 algV, 109 AlgW, 64, 117, 126, 127, 145 AlgX, 55 algX, 62 algZ, 119 Amino acids, 215 A-modules, 101 AmrZ, 117, 122 Amylase, 106 Amyloglucosidase, 198 Anaphylaxis, 230 Anti-inflammatory drug, 180, 229 Antireflux, 212 261
262 Arabidopsis thaliana, 146 Arc-Like DNA-Binding Protein, 122 Arthritis, 242 Ascophyllum nodosum, 2 Ascorbic acid, 192 Aspergillus niger, 192 Atomic force microscopy, 105 ATPase, 84 ATP-binding cassette importer, 73 Autoimmune, 247 Autophosphoryation, 121 Azadirachta indica, 198 Azomonas, 157 Azospirillum brasilense, 199 Azotobacter A. chrococcum, 159 A. vinelandii, 2
B Bacteriophage, 156 β-Barrel, 81 Beads, 177 micro, 175 nano, 175 Beijerinchia, 157 Bifidobacteria, 90 Bioartificial pancreas, 182, 183 Bioavailability, 220 Biocompatibility, 175, 195 Biocomposite, 187 Bioconversion, 193 Biodegradability, 175, 195 Biodegradation, 20 Biofilm(s), 120, 124 biofilm matrix, 135 Biofuel, 75 Biomaterial, 20, 107, 181 Biomedicine, 175 Biopolymers, 198 Bioreactor, 37, 191 Bioreduction, 188 Bioremediation, 75 Biosensor, 201 Biosorption, 198 Biotechnology, 107 Biotin, 201 Blends, 175, 177 Bovine serum albumin, 195
C Candida antarctica, 189 Candida guilliermondii, 193
Index Candida rugosa, 189 Candida tropicalis, 200 Capsules, 177 Carageenan, 220 Carbazole method, 99 Carbodiimide, 201 Carrageenan, 30, 189 Cartilage, 182 c-di-GMP, 59, 63, 123, 139 Cell-cell attachment, 109 Cellulose, 61, 63, 96, 106, 164, 220 C-5 epimerases, 8 Channel, 66 CheA, 121 Chitin, 63 Chitosan, 37, 106, 177, 198 Chlorella vulgaris, 199 Cholesterol, 221 Chondrocyte, 203 Chymotrypsin, 193 Citrate, 81 Citric acid cycle, 55 Clostridia, 219 Coalescence, 215 Coatings, 217 Coimmobilization, 199 Colitis, 231 Compression, 39 Computer-modeling, 15 Confocal microscopy, 33 Cross-linking, 177 Cyst, 96, 102 Cystic fibrosis, 74 Cytokine, 231 Cytotoxicity, 255
D D-cycloserine, 126 Degranulation, 230 DegS, 126 Denitrification, 199 Dental impressions, 212 Depolymerization, 19 Dermatan sulfate, 106 Derxia, 157 Desiccation, 141 Dexamethasone, 257 Dextran, 166 D-galactose, 191 Diclofenac, 179, 257 Diffusion, 45 Dilatometric, 29 Disaccharide, 86
Index DNA microarray, 86 Drug slow release, 175 Durvillaea antarctica, 2 Dye industry, 212
E Ecklonia maxima, 2 Ecosystem, 136 Edema, 242 eexDEF, 100 Egg-box model, 28, 29 Elasticity, 37 Electrodispersion, 185 Electroneutrality, 34, 36 Electron microscopy, 161 Electrostatic interactions, 12, 15, 29 β-Elimination, 18, 74, 107 Elongation, 39 Emulsifiers, 211 Emulsion, 215 Encapsulation, 46, 215 Endolyases, 20 Energy-at-break, 37 Enterobacteriaceae, 219 Enterobactin, 81 Enterochelin, 81 Enthalpy, 22 Enzyme-substrate complex, 105 Epimerase, 10 Epimerization, 109 Epiphytic, 145 Epithelia, 144 Erythema, 242 Escherichia coli, 8 Esterification, 212 Exopolysaccharide, 136, 141 Experimental autoimmune encephalomyelitis (EAE), 244–245
F Fermentation, 167 Ferrichrome, 81 Fertilizers, 90 Fibres, 177 Fibroblasts, 200 Fibrosarcoma, 255 Films, 177 First-order kinetics, 18 Flagellin, 79 Flexibility, 17 Flocculation, 215
263 Flow-cell systems, 141 Fructose 6-phosphate, 55 Furonan, 221
G α-Galactosidase, 183 α-D-Galacturonic acid, 26 Gastric fluid, 179 Gastrointestinal tract, 179 G-blocks, 4, 41 GDP-mannose, 60 GDP-mannose dehydrogenase, 55 Gelling, 10 Genomic, 125 Glomerulonephritis, 237, 252 Glucomannan, 107, 198 Gluconacetobacter xylinus, 61 Gluconeogenesis, 60 δ-Gluconolactone (GDL), 32 Glucuronan, 106 Glucuronan C-5-epimerase, 106 β-D-glucuronic acid, 107 Glucuronoxylohamnan, 221 Glutaraldehyde, 181 Glycosidic bonds, 26 Glycosphingolipids (GSL), 75 Glycosyltransferase, 55 β-Glycosyltransferases, 63 Grafting, 37 Guar gum, 181 Guluronic acid, 21 α-L-Guluronic acid, 4, 26, 56
H Heavy metal, 198 Heparan sulfate, 106 Heparin, 106 Herbicide, 198 Histidine decarboxylase, 230 Histidine kinase, 121 Homopolysaccharides, 98 Host-pathogen interactions, 147 Hyaluronate, 182 Hyaluronidase, 230 Hydrocolloid, 214 Hydrogel, 3, 20, 27, 31, 33, 35, 40, 187 Hydrogen bonding, 30 Hydrogen peroxide, 146 Hydrolysis, 99 Hydroxyapatite, 182 Hydroxyethylcellulose (HEC), 177
264 Hypercellularity, 240 Hyperemia, 231 Hypersensitive response (HR), 145 Hypochlorite, 144
I IL-6, 231 Imipenem, 145 Immobilization, 3, 33, 175, 185, 191, 215 Immune system, 3 Immunization, 240 Immunofluorescence, 238 Immunogenicity, 95 Immunoglobulin, 194 Immunogold electron microscopy, 76 Immunoisolation, 230 Inflammatory bowel disease (IBD), 231 Insulin, 198, 215 Interferon-γ, 144 Interleukin-6, 231 Ionic strength, 12, 14, 21
K Keratin, 201 Ketodeoxyphosphogluconate pathway, 57 Klebsiella aerogenes, 108
L Lactobacillus, 189 L. bifermentans, 193 L. bulgaricus, 200 Laminaria digitata, 2 Laminaria hyperborea, 2 Laminaria japonica, 2 Langerhans islets, 46 L-arabinose isomerase, 191 Large-scale production, 164 Laxation, 218 Lectin, 183 Lessonia nigrescens, 2 Leukocytes, 124 Leukotriene B4, 233 Levan, 164 Light-scattering, 15 Lipoprotein, 65 Low-angle laser light scattering, 25
M Macroalgae, 2 Macrocystis pyrifera, 2
Index Macrophage, 144 Magnetic resonance imaging, 33 Mannose 6-phosphate, 60 Mannuronan, 8, 99, 188 mannuronan C-5 epimerases, 17 Mannuronan C-5-epimerase, 98, 100 Mannuronic acid, 4, 21 GDP-mannuronic acid, 55 β−D-Mannuronic acid, 4, 56 Marginalan, 164 Mark-Houvink-Sakurada (MHS), 18 Markov model, 5 Matrix metalloproteinase type 2 (MMP-2), 255 M-blocks, 4 Membrane cytoplasmic membrane, 62 outer membrane, 62 Membrane science, 175 Methacrylate, 188 Metronidazole, 193 MG-block(s), 4, 41 Microaerophilic growth, 161 Microbeads, 183 Microencapsulation, 34, 185, 212, 215 Microflora, 216 Microsphere(s), 180, 201, 212, 230 Mitomycin C, 201 Monosaccharides, 90 Monte Carlo method, 15 MucA, 117, 127 MucB, 64, 124 MucD, 64 MucE, 64 Mucoid, 141 Mucoidy, 165 MucP, 64, 127 MucR, 64 Multiple sclerosis, 245 Multiprotein complex, 55 Mycobacterium M. flavum, 157 M. tuberculosis, 242
N N-(3-Aminopropyl)pyrrole, 200 Nanobeads, 183 Nephropathy, 248 Neutrophil, 144 Nicotiana tabacum, 191 Nitrilase, 188 Nitrogenase, 156, 157 Nitrogen-fixing bacteria, 161 Nonmucoid, 141
Index Nonmucoid, 120, 124 NtrC family, 121 Nuclear magnetic resonance, 6 Number-average molecular weight, 17
O O-acetylases, 95 O-acetylation, 140 Oligosaccharide, 74 Operon, 118 Organic acids, 215 Outer membrane, 56 Oxaloacetate, 55
P Pathoadaptive mucA, 117 Pectate, 106 Pectins, 96, 221 Pel, 141 Peptidoglycan, 125 Periplasm, 62, 73 Peroxide, 19 Phagocytosis, 118, 145 Pharmaceuticals, 31, 98 Phosphomannomutase, 55 Phosphomannose isomerase:GDP-mannose pyrophosphorylase, 55 pH sensitivity, 195 PilZ, 56, 139 Piroxicam, 256, 257 Planktonic, 120, 136 Plasma, 218 Pollution control, 175 Poly(ethylene glycol), 198 Poly(ethylene oxide), 182 Poly(guluronic acid), 106 Poly(hydroxybutyric acid), 156 Poly(L-lysine), 36 Poly(L-ornithine), 36 Poly(N-isopropylacrylamide) (PNIPAAm), 181 Poly(vinyl alcohol), 180 Polyacrylamide, 183 Polyanion, 12, 109 Polycation, 36, 46 Polydisperse, 17 Polyelectrolyte, 14, 15, 22, 35 Polyethyleneimine, 37 Poly(vinyl alcohol)-g-polyacrylamide, 179 Polymerization, 17, 18, 61, 63 Polyphenol oxidase, 201 Polyphenols, 19 Polysaccharide, 2
265 Polyuronan, 212 Porosity, 96 Porphyran, 221 Precursor, 55 Primers, 108 Probiotics, 216 Propylene glycol, 20 Propylene glycol alginates (PGAs), 212 Prostaglandin E2, 233 Proteases, 193 Proteomic, 125 Protozoa, 156 Pseudomonas P. aeruginosa, 2 P. alginovora, 108 P. delafieldii, 200 P. fluorescens, 8 P. methanitrificans, 157 P. putida, 8 P. syringae, 145, 162 P. syringae pv. Tomato, 146 Psl, 141 PsmE, 102 Pullulan, 106 Pyranose, 212
Q Quorum-sensing, 120
R γ-Radiation, 19 Raman spectroscopy, 33 Reactive oxygen, 146 Regulated intramembrane proteolysis, 125 Rheological properties, 167 Rhizobium meliloti, 106 Rhizosphere, 143 Rhl, 120 Rhodococcus erythropolis, 200 R-modules, 100 Rubber, 37
S σ22, 117 σ54, 121 Saccharomyces cerevisiae, 188 Satiety, 220 Seaweed, 2 Secretion, 61 Semisolid, 107
266 Shoe polish, 212 Shrinkage, 214 Silica, 185 Sphingomonads, 75 Sphingomonas sp. A1, 73 Sputum, 166 Stability, 38 Sterilization, 19 Stiffness, 37 Strain-hardening, 42 Streptococcus pyogenes, 60 Streptomyces psammoticus, 200 Stress-strain curve, 42 Subtilisin, 195 Sulfaminoheparosan, 106 Superchannel, 74, 78 Supercritical CO2, 179 Superoxide, 144 Swelling, 30 Syneresis, 34
T Tetrasaccharide, 86 Thickeners, 211 Thiobacillus sp., 199 TNF-α, 231 Tobacco, 146 TonB, 81 Transcriptome, 126 Transcriptomics, 86 Transmembrane domains, 83
Index Transplantation, 34, 230 Transporter, 81 Trisaccharide, 86 Twitching motility, 122 Type IV pilus, 122
U UDP-glucose, 62 UDP-glucose dehydrogenase, 60 Ulcerations, 231 Urease, 193 Urine, 218 Uronan, 106 Uronic acid(s), 4, 30
V Vegetative capsule, 102 Virulence factor, 117 Viscosity, 12, 14, 44
X Xanthan, 166 X-ray, 33 Xylan, 220 Xytitol, 193
Y Young’s modulus, 41